1: \documentclass[11pt,aps,showpacs]{revtex4}
2: %\documentclass[aps,twocolumn,showpacs]{revtex4}
3:
4: %\documentclass[aps,prb,preprint]{revtex4}
5:
6: \usepackage{epsfig}
7:
8: %\newcommand{\lbfig}[1]{\refstepcounter{fig} \label{#1} }
9: %\newcounter{fig}
10:
11: \newcommand{\nc}{\newcommand}
12: \nc{\be}{\begin{equation}}
13: \nc{\ee}{\end{equation}}
14: \nc{\bea}{\begin{eqnarray}}
15: \nc{\eea}{\end{eqnarray}}
16: %\nc{\bi}[1]{\bibitem{#1}}
17: \nc{\lsim}{\mbox{\raisebox{-.6ex}{~$\stackrel{<}{\sim}$~}}}
18: \nc{\gsim}{\mbox{\raisebox{-.6ex}{~$\stackrel{>}{\sim}$~}}}
19: \nc{\gtwid}{\mathrel{\raise.3ex\hbox{$>$\kern-.75em\lower1ex\hbox{$\sim$}}}}
20: \nc{\ltwid}{\mathrel{\raise.3ex\hbox{$<$\kern-.75em\lower1ex\hbox{$\sim$}}}}
21:
22: \begin{document}
23:
24: \vskip -0.2in
25:
26: \rightline{CERN-TH/2003-172, HD-THEP-03-38, UFIFT-HEP-03-20}
27:
28: \vskip 0.2in
29:
30: \title{Dynamics of super-horizon photons during inflation
31: with vacuum polarization}
32:
33: \author{Tomislav Prokopec}
34: \email{T.Prokopec@thphys.uni-heidelberg.de,
35: Tomislav.Prokopec@cern.ch
36: }
37: \affiliation{Institut f\"ur Theoretische Physik, Heidelberg
38: University,
39: Philosophenweg 16, D-69120 Heidelberg, Germany}
40:
41: \affiliation{Theory Division, CERN, CH-1211 Geneva 23, Switzerland}
42:
43: \author{Richard Woodard}
44: \email{woodard@phys.ufl.edu}
45: \affiliation{Department of Physics, University of Florida,
46: Gainesville, Florida 32611, USA}
47:
48: \date{\today}
49:
50:
51: \begin{abstract}
52:
53: We study asymptotic dynamics of photons propagating in the polarized vacuum
54: of a locally de Sitter Universe. The origin of the vacuum polarization is
55: fluctuations of a massless, minimally coupled, scalar, which we model
56: by the one-loop vacuum polarization tensor of scalar electrodynamics.
57: We show that late time dynamics of the electric field on superhorizon scales
58: approaches that of an Airy oscillator. The magnetic field amplitude,
59: on the other hand, asymptotically approaches a nonvanishing constant
60: (plus an exponentially small oscillatory component), which is suppressed
61: with respect to the initial (vacuum) amplitude. This implies that
62: the asymptotic photon dynamics is more intricate than that of a massive
63: photon obeying the local Proca equation.
64:
65: \end{abstract}
66:
67: \pacs{98.80.Cq, 04.62.+v}
68:
69: \maketitle
70:
71: \section{Introduction}
72: \label{Introduction}
73:
74: Even though we have recently witnessed significant progress in understanding
75: the dynamics of gauge fields (photons) during inflation, our understanding
76: remains quite rudimentary. In earlier work~\cite{ProkopecWoodard:2003,
77: ProkopecTornkvistWoodard:2002a,ProkopecTornkvistWoodard:2002b} we computed
78: the one loop vacuum polarization from charged, massless, minimally coupled
79: scalars in a locally de Sitter background. We also evaluated the integral
80: of the position-space vacuum polarization tensor against a tree order photon
81: wave function and found a nonzero result which grows with the same number of
82: scale factors as the mass term of the Proca equation.
83:
84: What we have so far done has a direct analogue in flat space quantum field
85: theory. Because the background possesses spacetime translation invariance
86: it is much simpler to work in momentum space. Gauge and Lorentz invariance
87: imply that the vacuum polarization is proportional to the transverse
88: projection operator, $\Pi^{\mu\nu}(p) = (p^2 \eta^{\mu\nu} - p^{\mu} p^{\nu})
89: \Pi(p^2)$. Here $\eta^{\mu\nu}$ is the Minkowski metric,
90: %
91: \begin{eqnarray}
92: \eta_{\mu\nu} = {\rm diag}(-1,1,1,1) \, , \qquad \eta^{\mu\nu} =
93: {\rm diag}(-1,1,1,1) \, ,\label{metric:eta}
94: \end{eqnarray}
95: %
96: and $p^2 \equiv p^{\mu} p^{\nu} \eta_{\mu\nu}$. One checks that the photon
97: remains massless by evaluating $p^2 \Pi(p^2)$ on the tree order mass shell,
98: $p^2 = 0$. A nonzero result --- as in the Schwinger model \cite{Schwinger:1962}
99: --- proves that quantum corrections change the tree order mass shell, but it
100: does not determine the fully corrected mass shell. For that one must find
101: the pole of the full propagator, that is, solve for $p^2$ such that,
102: \begin{equation}
103: p^2 \Bigl[1 - \Pi(p^2)\Bigr] = 0 \; . \label{fulleqn}
104: \end{equation}
105: What we have done for photons during inflation is the analogue of computing
106: $\Pi(p^2)$ at one loop and showing that $p^2 \Pi(p^2)\vert_{p^2 =0} \neq 0$.
107: What we do in this paper is the analogue of solving equation (\ref{fulleqn}).
108:
109: To introduce the problem, we remind the reader that we work in the conformal
110: coordinate system of a locally de Sitter space-time whose metric and inverse
111: are,
112: %
113: \begin{eqnarray}
114: g_{\mu\nu} = a^2 \eta_{\mu\nu} \,,\qquad g^{\mu\nu} = a^{-2} \eta^{\mu\nu}
115: \, . \label{metric}
116: \end{eqnarray}
117: %
118: Here $a$ is the scale factor,
119: %
120: \begin{equation}
121: a(\eta) = -\frac{1}{H\eta} \, , \qquad (\eta <0) \, . \label{scale factor}
122: \end{equation}
123: %
124: The Lagrangean density of massless minimally coupled scalar electrodynamics
125: in a general metric reads,
126: %
127: \begin{eqnarray}
128: {\cal L}_{\Phi QED} & = &
129: - \frac 14\sqrt{-g} g^{\mu\rho}g^{\nu\sigma}
130: F_{\mu\nu}F_{\rho\sigma}
131: - \sqrt{-g}g^{\mu\nu}
132: (\partial_{\mu} - ie A_{\mu}) \phi^*
133: (\partial_{\nu} + i e A_{\nu} ) \phi
134: \,,
135: \label{PhiQED}
136: \end{eqnarray}
137: %
138: where $F_{\mu\nu} \equiv \partial_{\mu} A_{\nu} - \partial_{\nu} A_{\mu}$.
139: Its reduction to the locally de Sitter background~(\ref{metric}--\ref{scale
140: factor}) is,
141: %
142: \begin{eqnarray}
143: {\cal L}_{\Phi QED} & \rightarrow &
144: - \frac 14 \eta^{\mu\rho}\eta^{\nu\sigma}
145: F_{\mu\nu}F_{\rho\sigma}
146: -a^2 \eta^{\mu\nu}
147: (\partial_{\mu} - ie A_{\mu}) \phi^*
148: (\partial_{\nu} + i e A_{\nu} ) \phi
149: \,,
150: \label{PhiQED:conformal}
151: \end{eqnarray}
152: %
153: where we made use of $\sqrt{-g} = a^4$. From this form of the Lagrangean,
154: it is obvious that the photon field couples conformally to gravity, with
155: the trivial rescaling,
156: %
157: \begin{eqnarray}
158: F_{\mu\nu} \rightarrow F_{\mu\nu}\qquad (A_\mu\rightarrow A_\mu)
159: \,.
160: \label{conformal-rescaling:photon}
161: \end{eqnarray}
162: %
163: An important consequence is that tree level photon wave functions take the
164: same form in conformal coordinates as they do in flat space, $A_\mu \propto
165: \epsilon_\mu {\rm e}^{ik\cdot x}$.
166:
167: On the other hand, the minimally coupled scalar field $\phi$
168: does not couple conformally to gravity, which can be easily seen
169: upon the conformal rescaling, $\phi \rightarrow a^{-1}\phi$ in
170: Eq.~(\ref{PhiQED:conformal}). This has as
171: a consequence the extensively studied particle creation during inflation,
172: known as superadiabatic amplification. At one loop this conformal symmetry
173: breaking is communicated to the photon sector by the cubic and quartic terms
174: in~(\ref{PhiQED:conformal}). The main objective of this work is to study how
175: this changes photon wave functions.
176:
177: Provided it couples minimally to gravity, an obvious candidate for
178: the scalar field $\phi$ is the charged component
179: of the fundamental Higgs scalar, which at low energies
180: manifests as the longitudinal component of the $W^{\pm}$ boson.
181: Of course, there may be many more charged scalars of this type
182: whose masses lie between that of the Higgs field, $m_H\sim 10^2$\,GeV,
183: and the energy scale of inflation, $H\simeq 10^{13}\,{\rm GeV}$. An important
184: class of such particles would be the supersymmetric scalar partners
185: of the standard model quarks and charged leptons.
186:
187: In the presence of charged scalar fluctuations the vacuum becomes polarized,
188: such that in conformal space times~(\ref{metric}) Maxwell's equations
189: generalize to~\cite{ProkopecWoodard:2003},
190: %
191: \begin{equation}
192: \eta^{\mu\nu} \eta^{\rho\sigma} \partial_{\rho} F_{\sigma\nu}(x) \!
193: + \!\!
194: \int \! d^4x' [\mbox{}^{\mu} \Pi_{\rm ret}^{\nu}]\!(x,x') A_{\nu}(x') \! = \! 0
195: % \eta^{\mu\nu}J_\nu^{\rm free}(x)
196: \, . \label{genMax}
197: \end{equation}
198: %
199: Here $[\mbox{}^{\mu} \Pi_{\rm ret}^{\nu}]\!(x,x')$ denotes the retarded
200: vacuum polarization bi-tensor, which has the general form,
201: %
202: \begin{eqnarray}
203: [\mbox{}^{\mu} \Pi_{\rm ret}^{\nu}]\!(x,x') &\equiv&
204: - [{}^\mu\! P^\nu] \, \chi_e(x,x')
205: + [{}^\mu\! \bar P^\nu]\, \delta n^2(x,x') \, . \quad \label{genPi}
206: \end{eqnarray}
207: %
208: The two transverse projectors are defined by,
209: %
210: \begin{equation}
211: [{}^\mu\! P^\nu] \equiv
212: [\eta^{\mu\nu} \eta^{\rho\sigma} - \eta^{\mu\rho} \eta^{\sigma\nu}]
213: \partial^{\prime}_{\rho}\partial_{\sigma}
214: \,,\qquad
215: [{}^\mu\! \bar P^\nu] \equiv
216: \eta^{\mu i} \eta^{\nu j} [\delta_{ij}\nabla^\prime\cdot\nabla
217: - \partial^\prime_i\partial_j]
218: \,,
219: \label{transverse:projectors}
220: \end{equation}
221: %
222: with $\partial_{\mu} \equiv \partial/\partial x^{\mu}$, $\partial_{\mu}^{
223: \prime} \equiv \partial/\partial x^{\prime\mu}$. The coefficient of
224: $[{}^\mu\! \bar P^\nu]$ in (\ref{genPi}) is,
225: %
226: \begin{equation}
227: \delta n^2(x,x') = \chi_e(x,x') + \frac{\chi_m}{1+\chi_m}(x,x')
228: \,.
229: \label{deltan2}
230: \end{equation}
231: %
232: where $\chi_e(x,x')$ and $\chi_m(x,x')$ denote
233: the electric and magnetic susceptibilities of the vacuum.
234: An important consistency check of equations~(\ref{genMax})--(\ref{genPi})
235: is obtained by taking a derivative $\partial_\mu$ of Eq.~(\ref{genMax}).
236: The first term vanishes by the antisymmetry of $F_{\sigma\nu}$, while
237: the integral term vanishes because of the transverse structure
238: of the vacuum polarization tensor~(\ref{genPi}),
239: %
240: \begin{equation}
241: \partial_\mu [{}^\mu\Pi^\nu] = 0 = \partial_\nu^\prime [{}^\mu\Pi^\nu]
242: \, . \label{dmu:muPinu}
243: \end{equation}
244: %
245: This is an immediate consequence of
246: %
247: $ \partial_\mu\; [{}^\mu\! P^\nu] = 0$,
248: $ \partial_\mu\;[{}^\mu\! \bar P^\nu] = 0$,
249: $ \partial^\prime_\nu [{}^\mu P^\nu] = 0$,
250: and $ \partial^\prime_\nu [{}^\mu\bar P^\nu] = 0$.
251:
252: \medskip
253:
254: In a locally de Sitter space-time~(\ref{scale factor})
255: there is a simple relation between ${\ell}(x;x')$, the
256: invariant length from $x^{\mu}$ to $x^{\prime \mu}$, and the conformal
257: coordinate interval ${\Delta x}^2(x;x')$,
258: \begin{equation}
259: 4 \sin^2\left(\frac12 H \ell(x;x') \right) = a a' H^2 {\Delta x}^2(x;x')
260: \equiv y(x;x') \; .
261: \end{equation}
262: We refer to $y(x;x')$ as the de Sitter length function. The scalar
263: propagator can be expressed in terms of $y \equiv y(x;x')$ and the two scale
264: factors,
265: %
266: \begin{eqnarray}
267: \lefteqn{{i \Delta}(x;x') = \frac{H^{D-2}}{(4 \pi)^{\frac{D}2}} \left\{ 2^{D-4}
268: \Gamma(D-1) \ln(a a') - \pi {\rm cot}(\pi {\scriptstyle \frac{D}2})
269: \frac{\Gamma(D-1)}{\Gamma(\frac{D}2)} \right.}
270: \nonumber \\
271: & & \hspace{3cm} \left. + \sum_{n=1}^{\infty} \biggl[ \frac1{n} \frac{
272: \Gamma(D-1+n)}{\Gamma(\frac{D}2 + n)} \Bigl(\frac{y}4\Bigr)^n - \frac1{n-
273: \frac{D}2} \frac{\Gamma(\frac{D}2 - 1 + n)}{\Gamma(n)} \Bigl(\frac{y}4\Bigr)^{
274: n - \frac{D}2}\biggr] \right\} \; .
275: \label{Delta}
276: \end{eqnarray}
277: %
278: Note that the homogeneous terms (i.e., $y^n$) and the constant terms have
279: slightly different proportionality constants from our previous
280: expression~\cite{OnemliWoodard:2002,ProkopecTornkvistWoodard:2002b}.
281: This has been done to make (\ref{Delta}) valid for regulating
282: a spacetime in which $D$ will ultimately be taken to some dimension other
283: than four~\footnote{We thank Ewald Puchwein for pointing out the
284: improvement.}. In $D=4$, the propagator~(\ref{Delta}) reduces to,
285: %
286: \begin{equation}
287: i \Delta(x;x') \; \stackrel{D\rightarrow 4}{\longrightarrow} \;
288: \frac{H^2}{4\pi^2}\bigg\{
289: \frac{\eta\eta'}{\Delta x^2}
290: - \frac 12 \ln(H^2\Delta x^2)
291: -\frac 14 + \ln(2)
292: \bigg\}
293: \,.
294: \label{Delta:4dim}
295: \end{equation}
296: %
297:
298: In the presence of charged scalar fluctuations described by this propagator,
299: the photon in a locally de Sitter space-time sees a polarized vacuum. This
300: effect is characterized by the renormalized vacuum polarization bi-tensor,
301: the one loop result for which is~\cite{ProkopecTornkvistWoodard:2002a,
302: ProkopecTornkvistWoodard:2002b},
303: %
304: \begin{eqnarray}
305: i \Bigl[\mbox{}^{\mu}\Pi^{\nu}_{\rm ren}\Bigr](x,x')
306: &=& \frac{\alpha}{32\pi^3}
307: \bigg\{
308: -\; [{}^\mu\!P^\nu]\;
309: \biggl[
310: \frac{1}{12}\partial^4\Big(\ln^2(\mu^2 {\Delta x}^2)
311: - 2\ln(\mu^2 {\Delta x}^2) \Big)
312: \nonumber \\
313: & & \hspace{2.6cm} +\; i\,{16 \pi^2 \over 3} \ln(a) \delta^4(x - x')
314: \nonumber\\
315: &&\hspace{2.6cm}
316: + \; \frac{1}{\eta\eta'}\partial^2\Big(\ln^2({H^\prime}^2 {\Delta x}^2)
317: + 2\ln({H^\prime}^2 {\Delta x}^2)
318: \Big)
319: \biggr]
320: \nonumber \\
321: & & \hspace{1.4cm}
322: +\; [{}^\mu\!\bar P^\nu]\;
323: \Bigl[ \frac{2}{\eta^2 \eta^{\prime 2}}
324: \Big(\ln^2({H^\prime}^2 {\Delta x}^2)
325: + 4\ln({H^\prime}^2 {\Delta x}^2)
326: \Big)
327: \Bigr]
328: \bigg\}
329: \, . \quad
330: \label{vacpol}
331: \end{eqnarray}
332: %
333: Here $\mu$ is the renormalization scale, $\alpha \equiv e^2/4\pi$ is the fine
334: structure constant, and $H^\prime \equiv 2^{-1}{\rm e}^{1/4} H$.
335: The polarization tensor~(\ref{vacpol}) was obtained using dimensional
336: regularization with the scalar propagator~(\ref{Delta}). It is fully
337: renormalized, which means that it is an integrable function of $x^{\prime
338: \mu}$ when the derivatives are taken after the integration. It is also gauge
339: invariant on account of its transversality on both indices~(\ref{dmu:muPinu}),
340:
341: The renormalized vacuum polarization (\ref{vacpol}) is properly an in-out
342: matrix element. To convert it into the retarded vacuum polarization
343: $[{}^\mu\Pi_{\rm ret}^\nu]$ of Eq.~(\ref{genMax}) we employ the
344: Schwinger-Keldysh formalism as has been described at length
345: elsewhere~\cite{ProkopecTornkvistWoodard:2002b}. The result takes the
346: same form as (\ref{genPi}) with electric susceptibility,
347: %
348: \begin{eqnarray}
349: \chi_e &=& \chi_e|_{\rm flat}
350: + \chi_e|_{\rm anomaly}
351: + \chi_e|_{\rm de\;Sitter}
352: \nonumber\\
353: \chi_e|_{\rm flat} &=& \frac{\alpha}{96\pi^2}\,\partial^4
354: \Big[
355: \Big(
356: \ln(\mu^2\Delta\tau^2)-1
357: \Big)
358: \theta(\Delta\eta)\theta(\Delta\tau^2)
359: \Big]
360: \nonumber\\
361: \chi_e|_{\rm anomaly} &=& \frac{\alpha}{6\pi}\,\delta^4(x-x'\,)
362: \nonumber\\
363: \chi_e|_{\rm de\;Sitter} &=& \frac{\alpha}{8\pi^2}H^2aa'\,\partial^2
364: \Big[
365: \Big(
366: \ln({H^\prime}^{\,2}\Delta\tau^2)+1
367: \Big)
368: \theta(\Delta\eta)\theta(\Delta\tau^2)
369: \Big]
370: \, .
371: \label{vacuum:pol:retarded:2}
372: \end{eqnarray}
373: %
374: The change in the index of refraction is,
375: %
376: \begin{eqnarray}
377: \quad\;
378: \delta n^2 &=& \frac{\alpha}{4\pi^2}H^4a^2{a'}^2\,
379: \Big(
380: \ln({H^\prime}^{\,2}\Delta\tau^2)+2
381: \Big)
382: \theta(\Delta\eta)\theta(\Delta\tau^2)
383: \,.
384: \label{vacuum:pol:retarded:3}
385: \end{eqnarray}
386: %
387: In these formulae $a\equiv a(\eta)$, $a'\equiv a(\eta')$, $H^\prime = 2^{-1}{
388: \rm e}^{1/4} H$, $\Delta \tau^2 = (\eta -\eta')^2 - \| \vec x-\vec x^{\,\prime}
389: \|^2$, $\Delta \eta = \eta-\eta'$, and $\theta(x) = 1$ for $x>0$, $\theta(x)
390: = 0$ for $x<0$.
391:
392: The electric susceptibility in~(\ref{vacuum:pol:retarded:2})
393: can be neatly split into three contributions.
394: The first contribution $\chi_e|_{\rm flat}$ is conformally invariant.
395: When expressed in terms of conformal time, this contribution becomes identical
396: to the one-loop electric susceptibility
397: in Minkowski (flat) space-time. This term contains no scale
398: factors, and it is renormalized precisely as in flat space.
399: The second contribution, $\chi_e|_{\rm anomaly}$, comes from the
400: trace anomaly, which arises as a consequence
401: of imperfect cancellation between the counterterm and
402: the local one-loop diagram in dimensional
403: regularization~\cite{ProkopecTornkvistWoodard:2002b}.
404: The anomalous contribution depends only weakly on
405: the scale factor $a$. The trace anomaly was first considered
406: in the context of electrodynamics in conformal space-times
407: in Ref.~\cite{Dolgov:1981}, while for a discussion
408: of scalar electrodynamics and nonabelian gauge theories see
409: Refs.~\cite{BirrellDavies:1982}
410: and~\cite{Dolgov:1993,ProkopecTornkvistWoodard:2002b}.
411: The third contribution to the electric susceptibility,
412: $\chi_e|_{\rm de\;Sitter}$ in~(\ref{vacuum:pol:retarded:2})
413: is purely inflationary, completely finite and grows linearly
414: with the scale factors $a$ and $a'$.
415: And finally, $\delta n^2$ in~(\ref{vacuum:pol:retarded:3})
416: is the second inflationary contribution, which
417: grows quadratically with the scale factors $a$ and $a'$.
418:
419: In section \ref{Photon dynamics with vacuum polarization in inflation}
420: we argue that only the one loop terms $\chi_{e\vert{\rm de\ Sitter}}$ and
421: $\delta n^2$ need be kept in equation (\ref{genMax}). We also give the
422: integro-differential equation which describes a spatial plane wave with
423: arbitrary time dependence. (The derivation is reserved for the Appendix.)
424: In section \ref{Late-time dynamics of superhorizon photons} we analyze
425: the integro-differential equation for superhorizon photons at
426: asymptotically late times. We show analytically that while the vector
427: potential approaches a nonzero constant, its time derivative behaves
428: like an Airy oscillator. Numerical analysis is also presented. In
429: section \ref{Discussion} we translate these results into statements
430: about the physical electric and magnetic fields.
431:
432: \section{Photon dynamics with vacuum polarization in inflation}
433: \label{Photon dynamics with vacuum polarization in inflation}
434:
435: We shall now study the dynamics of the photon field by including
436: only the genuinely new inflationary terms contributing
437: at one-loop to the vacuum polarization
438: tensor, $\chi_e|_{\rm de\;Sitter}$~(\ref{vacuum:pol:retarded:2})
439: and $\delta n$~(\ref{vacuum:pol:retarded:3}), since
440: their contribution to the vacuum polarization grows rapidly
441: with the scale factor. Indeed, $\delta n\propto a^2$,
442: and the electric susceptibility,
443: $\chi_e|_{\rm de\; Sitter}\propto a$. From
444: $\partial_{\sigma} a = \delta^0_{\sigma} H a^2$ it follows however,
445: that this term can also give an $a^2$ in Eq.~(\ref{genMax})
446: and has to, therefore, be treated on an equal footing as $\delta n^2$.
447: Moreover, we expect that, when compared with the one-loop contributions to
448: $\chi_e|_{\rm de\;Sitter}$ and $\delta n^2$, the contributions of higher loops
449: can be neglected, since they are down by more powers of $\alpha$,
450: but can give no extra powers of $a$.
451:
452: A general solution can be expressed as a linear combination of spatial
453: plane waves in Coulomb gauge,
454: %
455: \begin{equation}
456: A_{\mu}(x) = \epsilon_\mu(\vec k,\eta)\; {\rm e}^{i\vec k\cdot \vec x}
457: \,,\qquad
458: \epsilon_i k_i = 0
459: \,.
460: \label{vector:field:plane:wave}
461: \end{equation}
462: %
463: In the Appendix we show that $\epsilon_0(\vec{k},\eta)$ vanishes in the absence
464: of sources, just as it does in flat space. The Appendix also describes how
465: to perform the integral over $\vec {x}^{\,\prime}$ to obtain the following
466: integro-differential equation for the time dependent polarization vector,
467: %
468: \begin{eqnarray}
469: &&-(\partial_0^2 + k^2)\epsilon_i(\vec k,\eta)
470: + \frac{\alpha H^2}{\pi k}
471: \bigg[a(\partial_0^2 + {k}^{\,2})\!\int_{\eta_0}^\eta\! d\eta' a'
472: Y_1\Big(k\Delta\eta,\frac{k}{H^\prime}\Big)\,
473: \epsilon_i(\vec k,\eta^\prime)
474: \nonumber\\
475: &&\hspace{4.3cm}
476: +\; H^2a^2
477: \!\int_{\eta_0}^\eta\! d\eta' {a'}^2
478: %(\vec k,\eta')
479: \sin(k\Delta\eta)\,\epsilon_i(\vec k,\eta^\prime)
480: \bigg]
481: = 0
482: \, . \quad \label{eom:Ai}
483: \end{eqnarray}
484: %
485: Here $Y_1(x,\zeta)$ is,
486: %
487: \begin{equation}
488: Y_1(x,\zeta)
489: = \sin(x)\big[2\ln(2x/\zeta)+2
490: %\big]
491: + {\rm ci}(2x)-\gamma_E-\ln(2x)\big]
492: % + \sin(x)\big[{\rm ci}(2x)-\gamma_E-\ln(2x)\big]
493: - \cos(x)\big[{\rm si}(2x)+\pi/2\big]
494: \,.\quad
495: \label{Y1}
496: \end{equation}
497: %
498: Upon acting the time derivatives upon the first integral,
499: Eq.~(\ref{eom:Ai}) can be recast as,
500: %
501: \begin{eqnarray}
502: &&\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!
503: (\partial_0^2 + k^2)\epsilon_i(\vec k,\eta)
504: + a^2 \,\frac{2\alpha H^2}{\pi}
505: \bigg\{\bigg(\ln(a)
506: - \frac 54
507: % -\gamma_E + \ln\Big(\frac{H}{2\lambda}\Big)
508: \bigg)
509: \epsilon_i(\vec k,\eta)
510: - \frac{H^2}{k} \int_{\eta_0}^\eta d\eta' {a'}^2
511: \sin(k\Delta\eta)\epsilon_i(\vec k,\eta')
512: \nonumber\\
513: &&\hspace{3.5cm}
514: +\; \frac 1a \int_{\eta_0}^\eta\! d\eta' a'
515: \frac{1-\cos(k\Delta\eta)}{\Delta\eta}
516: \,\epsilon_i(\vec k,\eta^\prime)
517: - \int_{\eta_0}^\eta\! d\eta'
518: \ln\Big(1-\frac{\eta}{\eta'}\Big)
519: \partial_0^{\prime} \epsilon_i(\vec k,\eta^\prime)
520: \nonumber\\
521: &&\hspace{3.5cm}-\; \ln\Big(1-\frac 1a\Big)\epsilon_i(\vec k,\eta_0)
522: \bigg\}
523: = 0
524: .
525: \label{eom:Ai:2}
526: \end{eqnarray}
527: %
528:
529: This integral-differential equation is one of our main results. It should
530: describe photon dynamics accurately from the beginning of inflation, when
531: $a(\eta_0) = 1$ and the one loop terms are negligible, all the way to late
532: times, when $a \gg 1$ and the one loop terms become important. Another of
533: our main results is that, although certain aspects of this dynamical system
534: can be described in terms of local field equations, the system as a whole
535: is essentially nonlocal. There is solid physics behind this: the nonlocal
536: terms reflect the influence of scalar fluctuations within the past
537: light-cone. Note that this is perfectly consistent with causality. Note
538: further that, since $\ln(1-\eta/\eta')$ diverges logarithmically at $\eta'
539: = \eta$, the last integral cannot be recast in the form of an integral over
540: a kernel ${\tt K} = {\tt K}(\eta,\eta')$ multiplied by the polarization
541: vector, $\int_{\eta_0}^\eta{\tt K}(\eta,\eta') \epsilon_i(\vec k,\eta')$.
542: Instead, its form is that of a nonlocal conductivity, $\int_{\eta_0}^\eta{
543: \tt K'}(\eta,\eta') \partial_0^{\prime} \epsilon_i(\vec k,\eta')$. We shall
544: now address the relevance of these nonlocal (potentially dissipative)
545: contributions.
546:
547: \section{Late-time dynamics of superhorizon photons}
548: \label{Late-time dynamics of superhorizon photons}
549:
550: \subsection{Analytical analysis}
551:
552: We begin by performing a partial integration on Eq.~(\ref{eom:Ai:2}),
553: %
554: \begin{eqnarray}
555: &&(\partial_0^2 + k^2)\epsilon_i(\vec k,\eta)
556: + a^2H^2 \,\frac{2\alpha}{\pi}
557: \bigg\{
558: \bigg(
559: \ln\Big(\frac{k}{H}\Big)
560: - \frac 54 + \gamma_E
561: % + \ln\Big(\frac{H}{\lambda}\Big)
562: \bigg) \epsilon_i(\vec k,\eta)
563: \nonumber\\
564: &&\hspace{4.5cm}
565: -\; \int_{\eta_0}^\eta d\eta^\prime
566: \Big[
567: {\rm ci}(k\Delta\eta)+\frac{\sin(k\Delta\eta)}{k\eta^\prime}
568: \Big]\partial_0{^\prime}\epsilon_i(\vec k,\eta^\prime)\bigg\}
569: \nonumber\\
570: && \hspace{4.5cm} = \; a^2H^2 \,\frac{2\alpha}{\pi}
571: \bigg[{\rm ci}(k\Delta\eta_0)+\frac{\sin(k\Delta\eta_0)}{k\eta_0}\bigg]
572: \epsilon_i(\vec k,\eta_0)
573: \,.\qquad
574: \label{eom:IntDE:1}
575: \end{eqnarray}
576: %
577: It is now convenient to rewrite this equation in terms of the number of
578: e-foldings, $N \equiv \ln(a) = - \ln(-H\eta)$,
579: %
580: \begin{eqnarray}
581: &&\!\!\!\!\!\!\!\!\!
582: (\partial_N^2 + \partial_N + w^2a^{-2})\widetilde\epsilon_i(w,N)
583: + \frac{2\alpha}{\pi}
584: \bigg\{
585: \bigg(
586: \ln\big(w\big)
587: - \frac 54 + \gamma_E
588: \bigg)\widetilde\epsilon_i(w,N)
589: \nonumber\\
590: && \hspace{4.7cm}
591: -\; \int_{0}^N dN^\prime
592: \bigg[
593: {\rm ci}\big(w[{a^\prime}^{-1}-{a}^{-1}]\big)
594: -\frac{\sin\big(w[{a^\prime}^{-1}-{a}^{-1}]\big)}
595: {w{a^\prime}^{-1}}
596: \bigg]\partial_0{^\prime}\widetilde\epsilon_i(w,N^\prime)\bigg\}
597: \nonumber\\
598: &&\hspace{4.7cm} = \;\frac{2\alpha}{\pi}\bigg[
599: {\rm ci}\Big(w(1-1/a)\Big)
600: - \frac{\sin\Big(w(1-1/a)\Big)}{w}
601: \bigg]\widetilde\epsilon_i(w,0)
602: \, .\qquad
603: \label{eom:IntDE:2}
604: \end{eqnarray}
605: %
606: The new dependent variable is, $\widetilde\epsilon_i(w,N) \equiv
607: \epsilon_i(k,\eta)$, where $w \equiv k/H\gg 1$. The scale factors are $a =
608: {\rm e}^N$ and $a^\prime ={\rm e}^{N^\prime}$. When $N$ becomes large the
609: terms containing negative powers of $a$ can be dropped to give,
610: %
611: \begin{eqnarray}
612: && \!\!\!\!\!\!\!\!
613: (\partial_N^2 + \partial_N )\widetilde\epsilon_i(w,N)
614: + \frac{2\alpha}{\pi}
615: \bigg\{
616: \bigg(
617: \ln\big(w\big)
618: - \frac 54 + \gamma_E
619: \bigg) \widetilde\epsilon_i(w,N)
620: \nonumber\\
621: && \hspace{1.2cm}
622: - \int_{0}^N dN^\prime
623: \bigg[
624: {\rm ci}\big(w{a^\prime}^{-1}\big)
625: -\frac{\sin\big(w{a^\prime}^{-1}\big)}
626: {w{a^\prime}^{-1}}
627: \bigg]\partial_{N'}\widetilde\epsilon_i(w,N^\prime)\bigg\}
628: %&& \hspace{4.cm}
629: \simeq \;\frac{2\alpha}{\pi}\bigg[
630: {\rm ci}(w)
631: - \frac{\sin(w)}{w}
632: \bigg]\widetilde\epsilon_i(w,0)
633: \,.\qquad
634: \label{eom:IntDE:3}
635: \end{eqnarray}
636: %
637:
638: It is easy to see from equation (\ref{eom:IntDE:3}) that
639: $\widetilde\epsilon_i(w,N)$ approaches a constant plus an exponentially small
640: term. First note that the assumption is self consistent. If we make it then,
641: for large enough $N'$, the integrand of the nonlocal term falls off. This
642: means that the integral is independent of its upper limit and the constant
643: is determined by the relation,
644: \begin{equation}
645: \bigg(\!\! \ln\big(w\big) - \frac 54 + \gamma_E \!\!\bigg) \widetilde{
646: \epsilon}_i(w,\infty) - \! \int_{0}^{\infty} \!\!\!\! dN^\prime \bigg[
647: {\rm ci}\big(w{a^\prime}^{-1} \big) -\frac{\sin\big(w{a^\prime}^{-1}\big)}{
648: w{a^\prime}^{-1}} \bigg] \partial_{N'} \widetilde\epsilon_i(w,N^\prime)
649: \simeq \! \bigg[ {\rm ci}(w) - \frac{\sin(w)}{w} \bigg] \widetilde{
650: \epsilon}_i(w,0) \, .
651: \end{equation}
652: However, because the integral depends upon $\partial_{N'} \widetilde{
653: \epsilon}_i(w,N')$ for finite $N'$, when it is still significant, the actual
654: value of $\widetilde{\epsilon}_i(w,\infty)$ must be determined numerically.
655:
656: To get the rate at which $\widetilde{\epsilon}_i(w,N)$ approaches
657: $\widetilde{\epsilon}_i(w,\infty)$ we differentiate (\ref{eom:IntDE:3}) with
658: respect to $N$. Neglecting exponentially small terms gives the following
659: local equation for $\partial_N \widetilde{\epsilon}_i(w,N)$,
660: %
661: \begin{eqnarray}
662: \Big\{\partial_N^2 + \partial_N
663: + \frac{2\alpha}{\pi}
664: \Big(N - \frac 14\Big)\Big\}\partial_N\widetilde\epsilon_i(w,N)
665: \simeq 0
666: \, . \qquad
667: \label{eom:IntDE:5}
668: \end{eqnarray}
669: %
670: The general solution can be expressed in terms of Airy functions,
671: %
672: \begin{equation}
673: \partial_N\widetilde\epsilon_i(w,N)
674: \simeq a^{-1/2}\Bigl\{C_A\, {\rm Ai}\Big(-(2\alpha/\pi)^{1/3}
675: \Bigl[
676: N - \frac 14 - \frac{\pi}{8\alpha}
677: \Bigr]
678: \Big)
679: \,+\, C_B\, {\rm Bi}\Big(-(2\alpha/\pi)^{1/3}
680: \Bigl[
681: N - \frac 14 - \frac{\pi}{8\alpha}
682: \Bigr]
683: \Big)
684: \Bigr\}
685: \, . \; \label{Airy solution:2}
686: \end{equation}
687: %
688: Making use of the asymptotic expansion of the Airy functions~\footnote{The
689: asymptotic expansion for the Airy functions,
690: %
691: $$
692: {\rm Ai}(-z) \sim \frac{1}{\pi^{1/2}z^{1/4}}
693: \Big[
694: \cos\Big(\frac 23 z^{3/2} - \frac{\pi}{4}\Big)
695: + O(z^{-3/2})
696: \Big]\,,\qquad |z|\gg 1
697: $$
698: %
699: %
700: $$
701: {\rm Bi}(-z) \sim \frac{1}{\pi^{1/2}z^{1/4}}
702: \Big[
703: \sin\Big(\frac 23 z^{3/2} - \frac{\pi}{4}\Big)
704: + O(z^{-3/2})
705: \Big]\,,\qquad |z|\gg 1
706: \,.
707: \label{Airy functions:aasymptotic}
708: $$
709: %
710: },
711: Eq.~(\ref{Airy solution:2}) can be integrated to give asymptotically,
712: %
713: \begin{eqnarray}
714: \!\! \widetilde \epsilon_i(w,N) = \widetilde \epsilon_i(w,\infty)
715: + {\rm e}^{-N/2}
716: \Bigg\{\! \frac{C_A\sin\Big(\varphi[N,\alpha]\Big)
717: - C_B\cos\Big(\varphi[N,\alpha]\Big)}
718: {\pi^{1/2}(2\alpha/\pi)^{7/12}
719: \Big(N-\frac 14 - \frac{\pi}{8\alpha}\Big)^{3/4}}
720: + O(N^{-\frac 54})
721: \! \Bigg\}
722: \,, \;\; N\gg \ln(w)\,,
723: \quad
724: \label{Airy solution:epsiloni}
725: \end{eqnarray}
726: %
727: where,
728: %
729: \begin{equation}
730: \varphi[N,\alpha] \equiv \frac 23\Big[\frac{2\alpha}{\pi}\Big]^{1/2}
731: \Big[N-\frac 14 - \frac{\pi}{8\alpha}\Big]^{3/2}
732: - \frac\pi 4 \, .
733: \label{varphi:def}
734: \end{equation}
735: %
736: Of course the constants $C_A$ and $C_B$ are fixed by the behavior of
737: $\widetilde{\epsilon}_i(w,N)$ before the asymptotic form obtains. They
738: can only be determined numerically.
739:
740: \subsection{Numerical analysis}
741:
742: By performing a numerical study of Eq.~(\ref{eom:Ai:2}),
743: we now confirm the analytical results
744: of section~\ref{Late-time dynamics of superhorizon photons} above.
745: %
746: \begin{figure}[tb]
747: \vskip 0.1in
748: \leftline{\epsfig{file=phk40ExactvsInt1aB.eps,width=3.3in,height=3.2in }}
749: \vskip -3.2in
750: \rightline{\epsfig{file=phk40ExactvsInt1bB.eps, width=3.3in,height=3.2in }
751: \hskip 0.in}
752: \vskip -0.1in
753: \caption{
754: A comparison of the exact (numerical) solution of
755: the integro-differential equation~(\ref{eom:Ai:2}) with that of
756: Eqs.~(\ref{DE:zeta}-\ref{eom:DE+Int1}) for $k = 40 H$.
757: At late times the photon amplitude, $\epsilon_i$, approaches a constant.
758: The difference between the two solutions is smaller or of the order
759: the thickness of the lines, and quantitatively, it reaches about
760: 0.7\%.
761: \label{figure-I}
762: }
763: \end{figure}
764: %
765: The following approximation turns out to be quantitatively justified.
766: Upon neglecting the latter two integrals, Eq.~(\ref{eom:Ai:2})
767: reduces to,
768: %
769: \begin{eqnarray}
770: &&(\partial_0^2 + k^2)\epsilon_i(\vec k,\eta)
771: + a^2 \,\frac{2\alpha H^2}{\pi}
772: \Big\{\Big[\ln(a)
773: -\gamma_E + \ln\Big(\frac{H}{2\lambda}\Big)\Big]
774: \epsilon_i(\vec k,\eta)
775: - \zeta(\vec{k},\eta)
776: \Big\}
777: \simeq 0
778: \,,\qquad
779: \label{eom:DE+Int1}
780: \end{eqnarray}
781: %
782: where we define,
783: %
784: \begin{equation}
785: \zeta(\vec{k},\eta) \equiv \frac{H^2}{k} \int_{\eta_0}^\eta d\eta' {a'}^2
786: \sin(k\Delta\eta)\epsilon_i(\vec k,\eta')
787: \,,
788: \label{zeta}
789: \end{equation}
790: %
791: The function $\zeta(\vec{k},\eta)$ obeys the following differential equation
792: and initial conditions,
793: %
794: \begin{equation}
795: \Big(\frac{\partial^2}{\partial \eta^2} + k^2\Big) \zeta = a^2 H^2
796: \epsilon_i \qquad \zeta(\vec{k},\eta_0)=0
797: ,\quad \frac{\partial}{\partial \eta}\zeta(\vec{k},\eta)|_{\eta_0}=0
798: \,.
799: \label{DE:zeta}
800: \end{equation}
801: %
802:
803: To study how quantitative this approximation is, in figure~\ref{figure-I}
804: we compare the exact (numerical) solution of Eq.~(\ref{eom:Ai:2})
805: with the approximate one obtained by solving
806: Eqs.~(\ref{eom:DE+Int1}--\ref{DE:zeta}).
807: As can be seen in figure~\ref{figure-I}, the agreement is quite good. This
808: implies that the two final integrals in~(\ref{eom:Ai:2}) are dynamically
809: irrelevant.
810:
811: Figure~\ref{figure-I} shows that, at asymptotically late times,
812: $\widetilde \epsilon_i(w,N)$ is dominated by the constant value $\widetilde
813: \epsilon_i(w,\infty)$, whose real part is slightly suppressed with respect
814: to its classical asymptotic value of $\cos(40) \approx -.67$.
815: In figure~\ref{figure-III} we subtract the asymptotic constant
816: from the solution of Eqs.~(\ref{eom:DE+Int1}-\ref{DE:zeta}), multiply by
817: $e^{N/2}$, and compare the result with the Airy function. The figure
818: displays an oscillatory, slightly damped profile, in reasonably good
819: agreement with our result (\ref{Airy solution:epsiloni}).
820:
821: %
822: \begin{figure}[tbp]
823: \vskip -0.25in
824: \centerline{\epsfig{file=phk25DE2Proca.eps,width=4.3in}}
825: \caption{A comparison of the solution of~(\ref{eom:DE+Int1}--\ref{DE:zeta})
826: with that of the Airy oscillator~(\ref{eom:IntDE:5}).
827: }
828: \label{figure-III}
829: \vskip -0.1in
830: \end{figure}
831: %
832:
833: \section{Discussion}
834: \label{Discussion}
835:
836: It is illuminating to translate our results for the vector potential,
837: $A_i(x) = \epsilon_i(\vec{k},\eta) e^{i \vec{k} \cdot \vec{x}}$, into
838: statements about the behavior of the electric and magnetic fields. In
839: Coulomb gauge and using conformal coordinates, the densities of electric
840: and magnetic field lines per physical 3-volume are given by the following
841: expressions,
842: \begin{equation}
843: E^i(x) = \frac{\partial_0 A_i(x)}{a^2(\eta)} \qquad , \qquad B^i(x) =
844: -\frac{\epsilon^{ijk} \partial_j A_k(x)}{a^2(\eta)} \; .
845: \end{equation}
846: For spatial plane waves with $a(\eta) = e^N$ and $w = k/H$ we have,
847: \begin{equation}
848: e^{-i \vec{k} \cdot \vec{x}} E^i(x) = H e^{-N} \partial_N \widetilde{
849: \epsilon}_i(w,N) \qquad , \qquad e^{-i \vec{k} \cdot \vec{x}} B^i(x) =
850: -i \epsilon^{ijk} k^j e^{-2N} \widetilde{\epsilon}_k(w,N) \; .
851: \end{equation}
852: Keeping track only of powers of the scale factor, we can summarize the
853: results of section \ref{Late-time dynamics of superhorizon photons} by
854: saying that the full quantum system results in the following asymptotic
855: behavior,
856: \begin{equation}
857: \eta^{\mu\nu} \eta^{\rho\sigma} \partial_{\rho} F_{\sigma\nu}(x) \!
858: + \!\! \int \! d^4x' [\mbox{}^{\mu} \Pi_{\rm ret}^{\nu}]\!(x,x') A_{\nu}(x')
859: \! = \! 0 \quad \Longrightarrow \quad E^i \sim a^{-\frac32} \quad {\rm and}
860: \quad B^i \sim a^{-2} \; . \label{oursys}
861: \end{equation}
862:
863: It is interesting to contrast the asymptotic behavior (\ref{oursys}) of
864: the full, nonlocal system with that of various local models. The tree order
865: system gives,
866: \begin{equation}
867: \eta^{\mu\nu} \eta^{\rho\sigma} \partial_{\rho} F_{\sigma\nu} \! = \! 0
868: \quad \Longrightarrow \quad E^i \sim a^{-2} \quad {\rm and} \quad B^i
869: \sim a^{-2} \; . \label{tree}
870: \end{equation}
871: Although the behavior of the magnetic fields agree as regards powers of $a$,
872: we saw in Figure~\ref{figure-I} that the magnetic field of the actual
873: system analyzed in section III B is smaller by a factor of about 2/3. On the
874: other hand, the electric field of the actual system is enhanced by an
875: enormous factor of $a^{\frac12}$.
876:
877: Adding a Proca term with fixed mass $m_{\gamma} = g H$ results in the
878: following asymptotic behavior,
879: \begin{equation}
880: \eta^{\mu\nu} \Bigl(\eta^{\rho\sigma} \partial_{\rho} F_{\sigma\nu} -
881: g^2 H^2 a^2 A_{\nu} \Bigr) \! = \! 0 \quad \Longrightarrow \quad E^i
882: \sim a^{-\frac32 + \frac12 \sqrt{1 - 4 g^2}} \quad {\rm and} \quad B^i
883: \sim a^{-\frac52 + \frac12 \sqrt{1 - 4 g^2}} \; . \label{fixedm}
884: \end{equation}
885: The best perturbative fit to the actual system has $g^2 = 2 \alpha\ln(w)/\pi$
886: \cite{ProkopecTornkvistWoodard:2002a,ProkopecTornkvistWoodard:2002b} but
887: we see that the agreement between (\ref{oursys}) and (\ref{fixedm}) is
888: not good. The constant mass system has too much electric field and not
889: enough magnetic field.
890:
891: If the mass term grows as $m^2_{\gamma} = g^2 H^2 \ln(a)$ one finds,
892: \begin{equation}
893: \eta^{\mu\nu} \Bigl(\eta^{\rho\sigma} \partial_{\rho} F_{\sigma\nu} -
894: g^2 H^2 \ln(a) a^2 A_{\nu} \Bigr) \! = \! 0 \quad \Longrightarrow \quad E^i
895: \sim a^{-\frac32} \quad {\rm and} \quad B^i \sim a^{-\frac52} \; .
896: \label{mgrows}
897: \end{equation}
898: This model is suggested by mean field theory
899: \cite{DavisDimopoulosProkopecTornkvist:2000,
900: TornkvistDavisDimopoulosProkopec:2000,DimopoulosProkopecTornkvistDavis:2001}
901: with $g^2 = 2 \alpha/\pi$. It gives good agreement for the electric field
902: but results in too little magnetic field. In particular note that, whereas
903: the actual polarization vector approaches a nonzero constant, $\widetilde{
904: \epsilon}_i(w,\infty)$, the polarization vector of (\ref{mgrows}) approaches
905: zero.
906:
907: While none of the local models described above gives a perfect fit, the
908: effect of vacuum polarization is to vastly enhance the asymptotic electric
909: field with respect to its classical value in a manner most nearly described
910: by the growing mass of (\ref{mgrows}). Without regard to local analogues,
911: it is very apparent that vacuum polarization alters the kinematical
912: properties of superhorizon photons during inflation. On the other hand, the
913: fact that the asymptotic polarization vector is somewhat suppressed from
914: its classical value seems to indicate that there is no significant production
915: of photons.
916:
917: Exactly the opposite results seem to pertain for massless fermions which are
918: Yukawa coupled to a massless, minimally coupled
919: scalar~\cite{ProkopecWoodard:2003b}. In that case chirality conservation
920: protects the particle from gaining a mass, but the amplitude of the wave
921: function grows faster than exponentially during inflation. The physical
922: explanation seems to be copious particle production.
923:
924: \begin{acknowledgments}
925:
926: It is a pleasure to acknowledge Ola T\"ornkvist's collaboration
927: in much of the work reported here. We have also profited from
928: conversations with Dietrich B\"odeker, Anne-Christine Davis, Konstantinos
929: Dimopoulos, Sasha Dolgov, Hendrik J. Monkhorst, Glenn Starkman
930: and Tanmay Vachaspati. This work was partially supported by DOE contract
931: DE-FG02-97ER41029 and by the Institute for Fundamental Theory of
932: the University of Florida.
933:
934: \end{acknowledgments}
935:
936: \eject
937:
938: \section*{Appendix: Reduction of the polarization integral}
939: \label{Appendix: Reduction of the polarization integral}
940:
941: Here we show some details of the calculation of the inflationary
942: contribution to the vacuum polarization integral in~(\ref{genMax})
943: with the general plane wave (\ref{vector:field:plane:wave})
944: for the photon field,
945: $A_{\nu}(x') = \epsilon_\nu(\vec k,\eta')\; {\rm e}^{i\vec k\cdot \vec x'}$
946: $\;(k_i\epsilon_i = 0 = \epsilon_0)$.
947:
948: We begin our analysis by noting a useful identity,
949: %
950: \begin{eqnarray}
951: - [{}^\mu\!P^\nu]\; aa'
952: &=& -\, aa' \delta^\mu_0\delta^\nu_0 \nabla^2
953: + \bar\eta^{\mu\nu}[aa' (\nabla^2-\partial_0^2)
954: + H^2 a^2{a'}^2(1-\Delta\eta\partial_0)
955: ]
956: \nonumber\\
957: && -\, Ha{a'}^2 \delta_\mu^0\bar\partial^\nu
958: + Ha^2a' \bar\partial^\mu\delta^\nu_0
959: + aa'(\bar\partial^\mu\delta^\nu_0
960: + \delta^\mu_0\bar\partial^\nu)\partial_0
961: - aa' \bar\partial^\mu\bar\partial^\nu
962: \, .
963: \label{A:transverse-on:aa'}
964: \end{eqnarray}
965: %
966: In deriving this we made use of the relations,
967: %
968: \begin{equation}
969: \partial_\rho^{\,\prime} aa' = aa'\partial_\rho^{\,\prime}
970: + Ha{a'}^2 \delta_\rho^0
971: \qquad {\rm and} \qquad
972: \partial_\sigma aa' = aa'\partial_\sigma
973: + Ha^2a' \delta_\sigma^0
974: \,,
975: \label{A:partial:aa'}
976: \end{equation}
977: %
978: which imply,
979: %
980: \begin{equation}
981: \partial_\rho^{\,\prime}\partial_\sigma aa'
982: = H^2a^2{a'}^2 \delta_\rho^0\delta_\sigma^0
983: + Ha{a'}^2\delta_\rho^0\partial_\sigma
984: - Ha^2a'\partial^{\,\prime}_\rho\delta_\sigma^0
985: - aa'\partial^{\,\prime}_\rho\partial_\sigma
986: \, .
987: \label{A:partialpartial:aa'}
988: \end{equation}
989: %
990: Note also that we employ the notation,
991: %
992: \begin{equation}
993: \bar\partial^\mu \equiv \eta^{\mu i}\partial_i
994: \,,\qquad \bar \eta^{\mu\nu} \equiv \eta^{\mu\nu} + \delta^\mu_0\delta^\nu_0
995: \,.
996: \label{A:notation:eta+partial}
997: \end{equation}
998: %
999: When a derivative operator stands on the right of $aa'$, then it acts on a
1000: function of $x^{\mu} - x^{\prime \mu}$ and we can write $\partial_\rho^{\,
1001: \prime} = - \partial_\rho$. Moreover, due to the (spatial) transversality
1002: of the photon field, $\bar\partial^\nu A_\nu(x) = 0$, the terms
1003: in~(\ref{A:transverse-on:aa'}) containing $\bar\partial^\nu$ vanish
1004: identically.
1005:
1006: Consider first the $0$-th component of the photon field
1007: equation~(\ref{genMax}),
1008: %
1009: \begin{equation}
1010: - \nabla^2 A_0 - \frac{\alpha H^2}{8\pi^2}a \nabla^2 \partial^2
1011: \int d^4 x' \theta(\Delta\eta)\theta(\Delta \tau^2)
1012: \Bigl[
1013: \ln({H^\prime}^{\,2}\Delta\tau^2) + 1
1014: \Bigr] a' A_0(x') = 0
1015: \, ,
1016: \label{A:A0equation}
1017: \end{equation}
1018: %
1019: where we made use of
1020: Eqs.~(\ref{vacuum:pol:retarded:2}-\ref{vacuum:pol:retarded:3})
1021: and~(\ref{A:transverse-on:aa'}), and of
1022: the Coulomb gauge condition, $\nabla\cdot \vec A = 0$.
1023: Note that only the first term in~(\ref{A:transverse-on:aa'}) contributes
1024: to~(\ref{A:A0equation}). Furthermore, due to the spatial transverse
1025: structure of the Lorentz breaking term, $[{}^\mu\!\bar P^\nu]\delta n^2$,
1026: it does not contribute to~(\ref{A:A0equation}).
1027:
1028: At $\eta = \eta_0$ the integral drops out of (\ref{A:A0equation}) and we
1029: conclude that $\nabla^2 A_0(\vec{x},\eta_0) = 0$. We can make use of
1030: residual gauge freedom to make the unique solution be $A_0(\vec{x},\eta_0)
1031: = 0$. The same thing can be done for the first conformal time derivative.
1032: But then the equation implies that $A_0(\vec{x},\eta)$ vanishes for all
1033: $\eta$. Therefore, it is consistent to assume $\epsilon_0(\vec k,\eta) = 0$.
1034: This discussion also implies that the terms in~(\ref{A:transverse-on:aa'})
1035: which contain $\delta_\nu^0$ cannot contribute to the photon dynamical
1036: equation~(\ref{genMax}).
1037:
1038: Consider now the spatial components of equation~(\ref{genMax}).
1039: We start with the de Sitter contribution to the electric susceptibility,
1040: $\chi_e|_{\rm de\; Sitter}$~(\ref{vacuum:pol:retarded:2}),
1041: which with the help of~(\ref{A:transverse-on:aa'})
1042: (only the terms proportional to $\bar\eta^{\mu\nu}$ contribute),
1043: %
1044: \begin{eqnarray}
1045: &&-\eta_{i\mu}
1046: \int\! d^4x' [{}^\mu\!P^\nu] \chi_e(x,x')|_{\rm de\;Sitter}\;A_{\nu}(x')
1047: \;=\; \frac{\alpha H^2}{2\pi k^3}a {\rm e}^{i\vec k\cdot\vec x}
1048: (\partial_0^2 + {k}^{\,2})^2\int_{\eta_0}^\eta d\eta' a'
1049: \Xi_1\Big(k\Delta\eta,\frac{k}{H^\prime}\Big)\epsilon_i(\vec k,\eta')
1050: \nonumber\\
1051: &&\hspace{4.1cm}
1052: -\, \frac{\alpha H^4}{2\pi k^3}a^2 {\rm e}^{i\vec k\cdot\vec x}
1053: (\partial_0^2 + {k}^{\,2})\!\int_{\eta_0}^\eta d\eta'
1054: {a'}^2(1-\Delta\eta\partial_0)
1055: \Xi_1\Big(k\Delta\eta,\frac{k}{H^\prime}\Big)\epsilon_i(\vec k,\eta')
1056: \, . \qquad
1057: \label{A:chi_e:deSitter}
1058: \end{eqnarray}
1059: %
1060: Here $\eta_0 = -1/H$ denotes the conformal time at the beginning of
1061: inflation, at which $a(\eta_0) = 1$. We define $k \equiv \Vert \vec k \Vert$,
1062: $\Delta \eta = \eta - \eta '$. We also made use of the following elementary
1063: integrals,
1064: %
1065: \begin{eqnarray}
1066: && \int_0^{\Delta\eta} d^3 \Delta x\, {\rm e}^{-i\vec k\cdot \Delta \vec x}
1067: \Big\{
1068: \ln\Big[{H^\prime}^{\,2}(\Delta\eta^2-\Vert\Delta\vec x\Vert^2)
1069: \Big]
1070: + 1
1071: \Big\}
1072: \nonumber\\
1073: &&\hspace{4.cm}
1074: = 4\pi \int_0^{\Delta\eta} r^2 dr \frac{\sin(kr)}{kr}
1075: \Big\{
1076: \ln\Big[{H^\prime}^{\,2}(\Delta\eta^2-\Vert\Delta\vec x\Vert^2)
1077: \Big]
1078: + 1
1079: \Big\}
1080: \nonumber\\
1081: &&\hspace{4.cm}
1082: =\frac{4\pi}{k}\Delta\eta^2 \int_0^{1} x dx \sin(k\Delta\eta x)
1083: \Big\{
1084: \ln(1-x^2) + 2\ln({H^\prime}\Delta\eta)+1
1085: \Big\}
1086: \nonumber\\
1087: &&\hspace{4.cm}
1088: =\frac{4\pi}{k^3}\Xi_1(k\Delta\eta,k/{H^\prime})
1089: \,,
1090: \quad
1091: \end{eqnarray}
1092: %
1093: where $\Delta\vec x = \vec x - \vec x^{\,\prime}$,
1094: %
1095: \begin{eqnarray}
1096: \Xi_n(z,\zeta) &\equiv & [\sin(z)-x\cos(z)]
1097: [2\ln(z/\zeta)+n] + z^2\xi(z)
1098: \nonumber\\
1099: z^2\xi(z) &\equiv& z^2\int_0^1 z'dz' \sin(zz')\ln(1-{z'}^2)
1100: \nonumber\\
1101: &=& 2\sin(z)
1102: - [\cos(z)+z\sin(z)][{\rm si}(2z)+\pi/2]
1103: + [\sin(z)-z\cos(z)][{\rm ci}(2z)-\gamma_E-\ln(z/2)]
1104: \nonumber\\
1105: {\rm si}(z) &\equiv & -\int_z^\infty\frac{\sin(t)}{t}dt
1106: = \int_0^z\frac{\sin(t)}{t}dt - \frac{\pi}{2}
1107: \nonumber\\
1108: {\rm ci}(z) &\equiv & -\int_z^\infty\frac{\cos(t)}{t}dt
1109: = \int_0^z\frac{\cos(t)-1}{t}dt + \gamma_E + \ln(z)
1110: \,,
1111: \end{eqnarray}
1112: %
1113: and $\gamma_E \equiv - \psi(1) = -(d/dz)\ln[\Gamma(z)]|_{z=1} =
1114: 0.577\, 215\, 664.. $ is Euler's constant.
1115: Now making use of,
1116: %
1117: \begin{eqnarray}
1118: Y_1(z,\zeta)&\equiv& \frac 12 (\partial_z^2 + 1)\Xi_1(z,\zeta)
1119: \label{A:Y1}
1120: \\
1121: &=& \sin(z)\big[2\ln(2z/\zeta)+2
1122: + {\rm ci}(2z)-\gamma_E-\ln(2z)\big]
1123: - \cos(z)\big[{\rm si}(2z)+\pi/2\big]
1124: \nonumber\\
1125: \!\!\!\!
1126: (z\partial_z - 1)(\partial_z^2 + 1)\Xi_1(z,\zeta)
1127: &=& 4\sin(z) - 2\Xi_2(z,\zeta)
1128: \,,
1129: \end{eqnarray}
1130: %
1131: the integral~(\ref{A:chi_e:deSitter}) reduces to
1132: %
1133: \begin{eqnarray}
1134: &&\!\!\!\!\!\!
1135: -\eta_{i\mu}\int\! d^4x' [{}^\mu\!P^\nu] \chi_e(x,x')|_{\rm de\;Sitter}\;
1136: A_{\nu}(x')
1137: \;=\; \frac{\alpha H^2}{\pi k}a {\rm e}^{i\vec k\cdot\vec x}
1138: (\partial_0^2 + {k}^{\,2})\int_{\eta_0}^\eta d\eta' a'
1139: Y_1\Big(k\Delta\eta,\frac{k}{H^\prime}\Big)\epsilon_i(\vec k,\eta')
1140: \nonumber\\
1141: &&\hspace{5.cm}
1142: +\, \frac{\alpha H^4}{\pi k}a^2 {\rm e}^{i\vec k\cdot\vec x}
1143: \int_{\eta_0}^\eta d\eta' {a'}^2
1144: \Big[
1145: 2\sin(k\Delta\eta)-\Xi_2\Big(k\Delta\eta,\frac{k}{H^\prime}\Big)
1146: \Big]
1147: \epsilon_i(\vec k,\eta')
1148: \,.
1149: \qquad
1150: \label{A:chi_e:deSitter:2}
1151: \end{eqnarray}
1152: %
1153:
1154: The contribution from $\delta n^2$ to the spatial components
1155: of Eq.~(\ref{genMax}) can be obtained quite easily by the methods
1156: analogous to those employed for the contribution from
1157: $\chi_e|_{\rm de\; Sitter}$; the result can be written as,
1158: %
1159: \begin{eqnarray}
1160: \eta_{i\mu}\int\! d^4x' [{}^\mu\!\bar P^\nu] \delta n^2(x,x')\; A_{\nu}(x')
1161: \;=\; \frac{\alpha H^4}{\pi k}a^2 {\rm e}^{i\vec k\cdot\vec x}
1162: \int_{\eta_0}^\eta d\eta' {a'}^2
1163: \Xi_2\Big(k\Delta\eta,\frac{k}{H^\prime}\Big)\epsilon_i(\vec k,\eta')
1164: \,.
1165: \quad
1166: \label{A:delta n2}
1167: \end{eqnarray}
1168: %
1169: Remarkably, this contribution is fully canceled by the analogous
1170: one in~(\ref{A:chi_e:deSitter:2}). Keeping
1171: the terms that contribute at order $a^2$, we have,
1172: %
1173: \begin{eqnarray}
1174: \eta_{i\mu}\!\int\! d^4x' [{}^\mu\!\Pi_{\rm ret}^\nu](x,x')\, A_{\nu}(x')
1175: &\!=\!&\eta_{i\mu}\!\int\! d^4x'
1176: \Big(\!
1177: \!-\! [{}^\mu\!P^\nu] \chi_e(x,x')|_{\rm de\;Sitter}
1178: \!+\! [{}^\mu\!\bar P^\nu] \delta n^2(x,x')
1179: \Big) A_{\nu}(x')
1180: \!+\! O(a)
1181: \nonumber\\
1182: &\!=\!&\; \frac{\alpha H^2}{\pi k}a {\rm e}^{i\vec k\cdot\vec x}
1183: (\partial_0^2 + {k}^{\,2})\int_{\eta_0}^\eta d\eta' a'
1184: Y_1\Big(k\Delta\eta,\frac{k}{H^\prime}\Big)\epsilon_i(\vec k,\eta')
1185: \nonumber\\
1186: &\!+\!&
1187: %\hspace{5.9cm}
1188: \frac{2\alpha H^4}{\pi k}a^2 {\rm e}^{i\vec k\cdot\vec x}
1189: \int_{\eta_0}^\eta d\eta' {a'}^2
1190: \sin(k\Delta\eta)\epsilon_i(\vec k,\eta')
1191: + O(a)
1192: \,,
1193: \label{A:chi_e:deSitter+delta n2}
1194: \end{eqnarray}
1195: %
1196: where $O(a)$ indicates that the neglected terms can contribute
1197: to the vacuum polarization at most linearly in $a$.
1198:
1199: Now from the small argument expansions of the sine and cosine integral
1200: ({\it cf.} Eq.~[8.232] in Ref.~\cite{GradshteynRyzhik:1965}),
1201: %
1202: \begin{eqnarray}
1203: {\rm si}(z) &=& -\frac{\pi}{2} + \sum_{n=1}^\infty
1204: \frac{(-1)^{n+1}z^{2n-1}}{(2n-1)\,(2n-1)!}
1205: \,,\qquad |z|\ll 1
1206: \nonumber\\
1207: {\rm ci}(z) &=& \gamma_E + \ln(z) + \sum_{n=1}^\infty
1208: \frac{(-1)^{n}z^{2n}}{(2n)\,(2n)!}
1209: \,,\qquad\;\; |z|\ll 1
1210: \label{si:ci:small z}
1211: \end{eqnarray}
1212: %
1213: we obtain easily the following small argument expansion of $Y_1$,
1214: %
1215: \begin{eqnarray}
1216: Y_1(z,\zeta) &=& 2z\ln\Big(\frac{2z}{\zeta}\Big)
1217: - \frac 13 z^3\Big[
1218: \ln\Big(\frac{2z}{\zeta}\Big)
1219: - \frac{31}{12}
1220: \Big]
1221: + O(z^5\ln(z),z^5)
1222: \, .
1223: \label{A:Y1:expansion}
1224: \end{eqnarray}
1225: %
1226: This implies that, attempting to act with the derivative $\partial_0^2$
1227: on the integral in~(\ref{A:chi_e:deSitter+delta n2}), would result
1228: in a divergent contribution at the upper limit of integration, $\eta'=\eta$.
1229: In order to overcome this difficulty, observe first
1230: that one time derivative may be taken,
1231: %
1232: \begin{eqnarray}
1233: \frac{\partial_0}{k} \int_{\eta_0}^\eta d\eta' a'
1234: Y_1\Big(k\Delta\eta,\frac{k}{H^\prime}\Big)\epsilon_i(\vec k,\eta')
1235: = \int_{\eta_0}^\eta d\eta' a'
1236: \Big[\partial_z Y_1\Big(z,\frac{k}{H^\prime}\Big)\Big]_{z=k\Delta\eta}
1237: \epsilon_i(\vec k,\eta')
1238: \, .
1239: \end{eqnarray}
1240: %
1241: Note we can write,
1242: %
1243: \begin{eqnarray}
1244: \partial_z Y_1\big(z,\zeta\big)
1245: &=& \cos(z)\big[2\ln(2z/\zeta)+2+{\rm ci}(2z)-\gamma_E-\ln(2z)\big]
1246: + \sin(z)\big[{\rm si}(2z)+\pi/2\big]
1247: \nonumber\\
1248: &=& \Big[\partial_z Y_1\big(z,\zeta\big)-2\ln(2z/\zeta)-2\Big]
1249: + 2\ln(2z/\zeta) + 2
1250: \,,
1251: \end{eqnarray}
1252: %
1253: where in the last line we have added and subtracted the troublesome term
1254: which is logarithmically divergent at the upper limit of integration,
1255: $\eta'\rightarrow \eta$. We can now act with the second time derivative
1256: on the first term, to arrive at,
1257: %
1258: \begin{eqnarray}
1259: \frac{\partial^2_0+k^2}{k} \int_{\eta_0}^\eta\! d\eta' a'
1260: Y_1\Big(k\Delta\eta,\frac{k}{H^\prime}\Big)\epsilon_i(\vec k,\eta^\prime)
1261: &=& - 2\int_{\eta_0}^\eta\! d\eta' a'
1262: \frac{1-\cos(k\Delta\eta)}{\Delta\eta}
1263: \epsilon_i(\vec k,\eta^\prime)
1264: \nonumber\\
1265: &+& 2\partial_0\! \int_{\eta_0}^\eta
1266: d\eta' a'
1267: \Big[\ln(2H^\prime\Delta \eta)+1\Big]\epsilon_i(\vec k,\eta^\prime)
1268: \,.
1269: \qquad
1270: \label{Y1:integral}
1271: \end{eqnarray}
1272: %
1273: Now from,
1274: %
1275: \begin{eqnarray}
1276: a'\ln(2H^\prime\Delta\eta) &=& -\frac{1}{H\eta'}\Big[\ln(-2H^\prime\eta')
1277: - \sum_{n=1}^\infty\frac{1}{n}
1278: \Big(\frac{\eta}{\eta'}\Big)^n
1279: \Big]
1280: \nonumber\\
1281: &=& - \frac{1}{H}\frac{d}{d\eta'}
1282: \Big[
1283: \frac 12 \ln^2(-2H^\prime\eta')
1284: + \sum_{n=1}^\infty\frac{1}{n^2}\Big(\frac{\eta}{\eta'}\Big)^n\,
1285: \Big]
1286: \,,
1287: \nonumber
1288: \end{eqnarray}
1289: %
1290: we can rewrite the second integral in~(\ref{Y1:integral}) as follows,
1291: %
1292: \begin{eqnarray}
1293: \int_{\eta_0}^\eta\! d\eta' a'\ln(2H^\prime\Delta\eta)
1294: \epsilon_i(\vec k,\eta^\prime)
1295: &=& -\frac{1}{H}\bigg\{
1296: \Big[
1297: \frac 12 \ln^2(-2H^\prime\eta')
1298: + \sum_{n=1}^\infty\frac{1}{n^2}\Big(\frac{\eta}{\eta'}\Big)^n\,
1299: \Big]\epsilon_i(\vec k,\eta')
1300: \bigg\}\Big\vert_{\eta'=\eta_0}^{\eta'=\eta}
1301: \nonumber\\
1302: &+& \int_{\eta_0}^\eta\! d\eta'
1303: \Big[
1304: \frac 12 \ln^2(-2H^\prime\eta')
1305: + \sum_{n=1}^\infty\frac{1}{n^2}\Big(\frac{\eta}{\eta'}\Big)^n\,
1306: \Big]
1307: \frac{d}{d\eta'}\epsilon_i(\vec k,\eta^\prime)
1308: \, .
1309: \label{Y1:integral:3}
1310: \end{eqnarray}
1311: %
1312: Its time derivative equals,
1313: %
1314: \begin{eqnarray}
1315: \partial_0\int_{\eta_0}^\eta\! d\eta' a'
1316: \Big[
1317: \ln(2H^\prime\Delta\eta)+1
1318: \Big]
1319: \epsilon_i(\vec k,\eta^\prime)
1320: &=& a\Big[\ln(-2H^\prime\eta)+1\Big]\epsilon_i(\vec k,\eta)
1321: + a\ln\Big(1-\frac{\eta}{\eta_0}\Big)\epsilon_i(\vec k,\eta_0)
1322: \nonumber\\
1323: &+& a\int_{\eta_0}^\eta\! d\eta' \ln\Big(1-\frac{\eta}{\eta'}\Big)
1324: \frac{d}{d\eta'}\epsilon_i(\vec k,\eta^\prime)
1325: \,.
1326: \label{Y1:integral:4}
1327: \end{eqnarray}
1328: %
1329: Upon collecting all of the terms, Eq.~(\ref{A:chi_e:deSitter+delta n2})
1330: reduces to,
1331: %
1332: \begin{eqnarray}
1333: \eta_{i\mu}\!\int\! d^4x' [{}^\mu\!\Pi_{\rm ret}^\nu](x,x')\, A_{\nu}(x')
1334: &=& -\, a^2\frac{2\alpha H^2}{\pi}{\rm e}^{i\vec k\cdot\vec x}
1335: \bigg[-\Big(\ln(-2H^\prime\eta)+1\Big)\epsilon_i(\vec k,\eta)
1336: %\ln\Big(\frac{a}{2}\Big)\epsilon_i(\vec k,\eta)
1337: \nonumber\\
1338: &&-\;
1339: \frac{H^2}{k}
1340: \int_{\eta_0}^\eta d\eta' {a'}^2
1341: \sin(k\Delta\eta)\epsilon_i(\vec k,\eta')
1342: \nonumber\\
1343: &&+\; \frac 1a \int_{\eta_0}^\eta\! d\eta' a'
1344: \frac{1-\cos(k\Delta\eta)}{\Delta\eta}
1345: \epsilon_i(\vec k,\eta^\prime)
1346: \nonumber\\
1347: &&-\;
1348: \int_{\eta_0}^\eta\! d\eta'
1349: \ln\Big(1-\frac{\eta}{\eta'}\Big)
1350: \frac{d}{d\eta'}\epsilon_i(\vec k,\eta^\prime)
1351: \nonumber\\
1352: &&-\; \ln\Big(1-\frac{1}{a}\Big)\epsilon_i(\vec k,\eta_0)
1353: \bigg]
1354: \,,
1355: \label{A:chi_e:deSitter+delta n2:2}
1356: \end{eqnarray}
1357: %
1358: where we took account of $\eta_0 = -1/H$. When $a\gg 1$,
1359: the last local term, which depends on the initial photon amplitude
1360: $\epsilon_i(\vec k,\eta_0)$, contributes as $O(a^{-1})$, such that,
1361: in the limit when $a\gg 1$, it can be consistently neglected.
1362:
1363: \begin{thebibliography}{99}
1364:
1365: \bibitem{ProkopecWoodard:2003}
1366: T.~Prokopec and R.~P.~Woodard,
1367: ``Vacuum polarization and photon mass in inflation,''
1368: arXiv:astro-ph/0303358, accepted for publication
1369: by the {\it American Journal of Physics}.
1370: %%CITATION = ASTRO-PH 0303358;%%
1371:
1372: \bibitem{ProkopecTornkvistWoodard:2002a}
1373: T. Prokopec, O. T\"ornkvist and R. Woodard,
1374: ``Photon mass from inflation,''
1375: Phys. Rev. Lett. {\bf 89}, 101301-1--101301-5 (2002)
1376: [arXiv:astro-ph/0205331].
1377: %%CITATION = ASTRO-PH 0205331;%%
1378:
1379: \bibitem{ProkopecTornkvistWoodard:2002b}
1380: T. Prokopec, O. T\"ornkvist, and R. P. Woodard,
1381: ``One loop vacuum polarization in a locally de Sitter background,''
1382: Ann. Phys. {\bf 303}, 251--274 (2003) [arXiv:gr-qc/0205130].
1383: %%CITATION = GR-QC 0205130;%%
1384:
1385: \bibitem{Schwinger:1962}
1386: J. Schwinger,
1387: ``Gauge Invariance and Mass II,''
1388: Phys. Rev. {\bf 128}, 2425--2429 (1962).
1389: %%CITATION = PHRVA,128,2425;%%
1390:
1391: \bibitem{OnemliWoodard:2002}
1392: V.~K.~Onemli and R.~P.~Woodard,
1393: ``Super-acceleration from massless, minimally coupled phi**4,''
1394: Class.\ Quant.\ Grav.\ {\bf 19} (2002) 4607
1395: [arXiv:gr-qc/0204065].
1396: %%CITATION = GR-QC 0204065;%%
1397:
1398: \bibitem{Dolgov:1981}
1399: A.~D.~Dolgov,
1400: ``Conformal Anomaly And The Production Of Massless Particles
1401: By A Conformally Flat Metric,''
1402: Sov.\ Phys.\ JETP {\bf 54} (1981) 223
1403: [Zh.\ Eksp.\ Teor.\ Fiz.\ {\bf 81} (1981) 417].
1404: %%CITATION = SPHJA,54,223;%%
1405:
1406: \bibitem{BirrellDavies:1982}
1407: N.~D.~Birrell and P.~C.~Davies,
1408: ``Quantum Fields In Curved Space,'' Cambridge University Press, Cambridge,
1409: England (1982).\\
1410: F.~D.~Mazzitelli and F.~M.~Spedalieri,
1411: ``Scalar electrodynamics and primordial magnetic fields,''
1412: Phys.\ Rev.\ D {\bf 52} (1995) 6694
1413: [arXiv:astro-ph/9505140].
1414: %%CITATION = ASTRO-PH 9505140;%%
1415:
1416: \bibitem{Dolgov:1993}
1417: A.~Dolgov,
1418: ``Breaking of conformal invariance
1419: and electromagnetic field generation in the universe,''
1420: Phys.\ Rev.\ D {\bf 48} (1993) 2499
1421: [arXiv:hep-ph/9301280].
1422: %%CITATION = HEP-PH 9301280;%%.
1423:
1424: \bibitem{DavisDimopoulosProkopecTornkvist:2000}
1425: A.-C. Davis, K. Dimopoulos, T. Prokopec, and O. T\"orn\-kvist,
1426: ``Primordial spectrum of gauge fields from inflation,''
1427: Phys. Lett. B {\bf 501}, 165--172 (2001)
1428: [arXiv:astro-ph/0007214].
1429: %%CITATION = ASTRO-PH 0007214;%%
1430:
1431: \bibitem{TornkvistDavisDimopoulosProkopec:2000} O. T\"ornkvist,
1432: A.-C. Davis, K. Dimopoulos, and T. \-Pro\-ko\-pec, ``Large scale
1433: primordial magnetic fields from inflation and preheating,''
1434: in {\sl Verbier 2000, Cosmology and particle physics} (CAPP2000),
1435: 443--446 [arXiv:astro-ph/0011278].
1436: %%CITATION = ASTRO-PH 0011278;%%
1437:
1438: \bibitem{DimopoulosProkopecTornkvistDavis:2001} K. Dimopoulos, T.
1439: Prokopec, O. T\"ornkvist, and A.-C. Davis, ``Natural magnetogenesis
1440: from inflation,'' Phys. Rev. D {\bf 65}, 063505-1--063505-26
1441: (2002) [arXiv:astro-ph/0108093].
1442: %%CITATION = ASTRO-PH 0108093;%%
1443:
1444: \bibitem{ProkopecWoodard:2003b}
1445: T.~Prokopec and R.~P.~Woodard,
1446: ``Production of Massless Fermions during Inflation,''
1447: arXiv:astro-ph/0309593.
1448: %%CITATION = ASTRO-PH 0309593;%%
1449:
1450: \bibitem{GradshteynRyzhik:1965}
1451: {\tt GR:} Izrail Solomonovich Gradshteyn, Iosif Moiseevich Ryzhik,
1452: {\it Table of integrals, series, and products}, 4th edition,
1453: Academic Press, New York (1965).
1454:
1455: \end{thebibliography}
1456:
1457: \end{document}
1458: