hep-ph0109176/text
1: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2: % Fermion production from preheating-amplified metric perturbations
3: % 
4: % ST 10-13,19,25-27/8/01,29/8/01-4/9/01,8/9/01,14-16/9/01
5: % BB 10/8/01, 29/8/01, 18/9/01
6: % 
7: % MP,LS,ST,BB              
8: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
9: 
10: \documentstyle[prd,eqsecnum,aps,epsf]{revtex}
11: %\documentclass[draft]{JHEP}
12: %\usepackage{epsfig}
13: %\usepackage{amssymb}
14: %\usepackage{bbm}
15:  
16:  
17: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
18: \newcommand{\beq}{\begin{equation}}
19: \newcommand{\beqn}{\begin{eqnarray}} 
20: \newcommand{\eeq}{\end{equation}}
21: \newcommand{\eeqn}{\end{eqnarray}}
22: \newcommand{\beqa}{\begin{eqnarray}}
23: \newcommand{\eeqa}{\end{eqnarray}}
24: \newcommand{\mpl}{M_p}
25: \newcommand{\k}{{\kappa}}
26: \newcommand{\lmk}{\left(}
27: \newcommand{\rmk}{\right)}
28: \newcommand{\lkk}{\left[}
29: \newcommand{\rkk}{\right]}
30: \newcommand{\lnk}{\left\{}
31: \newcommand{\rnk}{\right\}}
32: \newcommand{\zk}{z_k}
33: \newcommand{\call}{{\cal L}}
34: \newcommand{\calr}{{\cal R}}
35: \newcommand{\half}{\frac{1}{2}}
36: \newcommand{\kc}{\kappa\chi}
37: \newcommand{\bkc}{\beta\kappa\chi}
38: \newcommand{\gkc}{\gamma\kappa\chi}
39: \newcommand{\gbkc}{(\gamma-\beta)\kappa\chi}
40: \newcommand{\dchi}{\delta\chi}
41: \newcommand{\dsigma}{\delta\sigma}
42: \newcommand{\dOmega}{\delta\Omega}
43: \newcommand{\Phibd}{\Phi_{\rm BD}}
44: \newcommand{\echi}{\epsilon_\chi}
45: \newcommand{\esigma}{\epsilon_\sigma}
46: \newcommand{\Phihat}{\hat{\Phi}}
47: \newcommand{\Psihat}{\hat{\Psi}}
48: \newcommand{\ahat}{\hat{a}}
49: \newcommand{\that}{\hat{t}}
50: \newcommand{\Hhat}{\hat{H}}
51: \newcommand{\ab}{q}
52: 
53: \newcommand{\sign}{\mathop{\rm sign}\nolimits}
54: \newcommand{\dett}{\mathop{\rm det}\nolimits}
55: \newcommand{\Ree}{\mathop{\rm Re}\nolimits}
56: \newcommand{\identity}{\mathbbm{1}}
57: \newcommand{\dbar}{\partial \hskip -5.5pt/ }
58: \newcommand{\gsim}{\mbox{\raisebox{-1.ex}{$\stackrel
59:      {\textstyle>}{\textstyle\sim}$}}}
60: \newcommand{\lsim}{\mbox{\raisebox{-1.ex}{$\stackrel
61:      {\textstyle<}{\textstyle \sim}$}}}
62: 
63: %\renewcommand{\theequation}{\thesection.\arabic{equation}}
64: \newcommand{\dddot}[1]{\stackrel{...}{#1}}
65: \newcommand{\pkt}{\; .}
66: \newcommand{\kma}{\; ,}
67: \newcommand{\tr}{{\rm tr \; }}
68: \newcommand{\Tr}{{\rm  Tr  \;}}
69: \newcommand{\re}[1]{{\rm Re }\,#1}
70: \newcommand{\im}[1]{{\rm Im }\,#1}
71: \newcommand{\pslash}{ \slash\hspace{-2.3mm} p}
72: \newcommand{\calF}{{\cal F}}
73: \newcommand{\calC}{{\cal C}}
74: \newcommand{\calS}{{\cal S}}
75: \newcommand{\calE}{{\cal E}}
76: \newcommand{\calL}{{\cal L}}
77: \newcommand{\intk}{\int\frac{d^3k}{(2\pi)^3 2\omega_{k0}}\,}
78: \newcommand{\intkp}{\int\frac{d^3k'}{(2\pi)^3 2\omega_{k'0}}\,}
79: \newcommand{\intp}{\int\frac{d^3p}{(2\pi)^3 2\Ep\,}}
80: \newcommand{\intpp}{\int\frac{d^3p'}{(2\pi)^3 2\Epp\,}}
81: \newcommand{\intppp}{\int\frac{d^3p''}{(2\pi)^3 2\Eppp\,}}
82: \newcommand{\bfk}{{\bf k}}
83: \newcommand{\bfp}{{\bf p}}
84: \newcommand{\bfq}{{\bf q}}
85: \newcommand{\bfx}{{\bf x}}
86: \newcommand{\bgamma}{\mbox{\boldmath{$\gamma$}\unboldmath}}
87: \newcommand{\bsigma}{\mbox{\boldmath{$\sigma$}\unboldmath}}
88: \newcommand{\bSigma}{\mbox{\boldmath{$\Sigma$}\unboldmath}}
89: \newcommand{\be}{\begin{equation}}
90: \newcommand{\ee}{\end{equation}}
91: \newcommand{\bea}{\begin{eqnarray}}
92: \newcommand{\eea}{\end{eqnarray}}
93: \newcommand{\omk}{\omega_{k0}}
94: \newcommand{\omkp}{\omega_{k'0}}
95: \newcommand{\Ep}{E_{p}}
96: \newcommand{\Epp}{E_{p'}}
97: \newcommand{\Eppp}{E_{p''}}
98: 
99: %%%%% singlefig %%%%%
100: \newcommand{\singlefig}[2]{
101: \begin{center}
102: \begin{minipage}{#1}
103: \epsfxsize=#1
104: \epsffile{#2}
105: \end{minipage}
106: \end{center}}
107: %
108: %%%%% figcaption %%%%%
109: \newenvironment{figcaption}[2]{
110:  \vspace{0.3cm}
111:  \refstepcounter{figure}
112:  \label{#1}
113:  \begin{center}
114:  \begin{minipage}{#2}
115:  \begingroup \small FIG. \thefigure: }{
116:  \endgroup
117:  \end{minipage}
118:  \end{center}}
119: %
120: 
121: \begin{document}
122: 
123: \title{Fermion production from preheating-amplified metric perturbations}
124: \author{Bruce A. Bassett$^1$, Marco Peloso$^2$, Lorenzo Sorbo$^3$
125: and Shinji Tsujikawa$^4$ }
126: \address{$^1$ Relativity and Cosmology Group (RCG), University of
127: Portsmouth, Mercantile House, Portsmouth,  PO1 2EG, 
128: England\\[.3em]}
129: \address{$^2$ Physikalisches Institut, Universit{\"{a}}t Bonn Nussallee 12,
130: D-53115 Bonn, Germany}
131: \address{$^3$ SISSA/ISAS, via Beirut 2-4, I-34013 Trieste, Italy, 
132: and INFN, Sezione di Trieste, via Valerio 2, I-34127 Trieste, Italy}
133: \address{$^4$ Research Center for the Early Universe, 
134: University of Tokyo, Hongo, Bunkyo-ku, Tokyo 113-0033, Japan} 
135: \date{\today} 
136:  \maketitle
137: \begin{abstract}
138: 
139: We study gravitational creation of light fermions in the presence of 
140: classical scalar metric perturbations about a flat Friedmann-Lemaitre-
141: Robertson-Walker (FLRW) background. These perturbations can be large
142: during preheating, breaking the conformal flatness of the background
143: spacetime. We compute numerically the total number of particles generated
144: by the modes of the metric perturbations which have grown sufficiently to
145: become classical. In the absence of inhomogeneities massless fermions are
146: not gravitationally produced, and then this effect may be relevant for
147: abundance estimates of light gravitational relics.
148: 
149: 
150: \end{abstract}
151: \vskip 1pc 
152: \pacs{pacs: 98.80.Cq}
153: \vskip 2pc
154: 
155: 
156: %%%%%%%%%%%%%%%%%%%%%%%%%%
157: \section{Introduction}                            
158: %%%%%%%%%%%%%%%%%%%%%%%%%%
159: 
160: 
161: It is well known that particles are created in an expanding Universe
162: \cite{BD80}. Pioneering work by Parker \cite{Parker} highlighted the
163: creation of nonconformally-invariant particles even in flat 
164: Friedmann-Lemaitre-Robertson-Walker (FLRW) Universes. The extension to
165: anisotropic Bianchi cosmologies by Zel'dovich and Starobinsky \cite{ZS},
166: showed that massless, conformally coupled, scalar particles are created
167: due to the breaking of conformal invariance which, in four dimensions, is
168: signaled by a non-vanishing Weyl tensor. In the presence of small,
169: inhomogeneous metric perturbations, Horowitz and Wald \cite{HW} obtained
170: the vacuum expectation value of the stress tensor for a conformally
171: invariant field and showed that particle production only occurs if
172: $\langle T_{\mu\nu}(x)\rangle$ is non-local (i.e., has non-zero
173: contributions with support on the past-null cone of the event $x$).
174: 
175: The number of particles created by inhomogeneities can be computed by a
176: perturbative evaluation of the $S$-matrix~\cite{Frieman,campos,CV}. In the
177: inhomogeneous case, the occupation number of produced particles is
178: composed by three parts, {\em viz}, the zeroth-order contribution due to
179: the homogeneous expansion, a first-order (in the  perturbation amplitude)
180: part arising from the interference between 0- and 2- particle states, and
181: a second-order contribution which comes from the interaction between
182: nonzero particle states. In the case of massive or nonconformally coupled
183: scalar fields, the last two terms (respectively linear and quadratic in
184: the metric perturbations) typically give a contribution small compared
185: with  the first one. However, in the massless and conformally coupled
186: cases,  the first two terms vanish in rigid, exactly FLRW backgrounds, and
187: only inhomogeneity contributes to the gravitational particle production.
188: 
189: Such theoretical studies were historically of interest both because they
190: focus on the interplay between quantum field theory and gravity and because
191: of their relevance to Misner's ``chaotic cosmology'' program \cite{misner}
192: in which arbitrary initial inhomogeneity and anisotropy were to be damped
193: to acceptable levels due to particle production. The non-renormalisability
194: of standard quantum gravity, the successes of string theory and the
195: dominance of cosmological inflation have since removed much of the
196: motivation for studies of particle production from inhomogeneity, though
197: research on non-equilibrium issues, generalized fluctuation-dissipation
198: relations and effective Einstein-Langevin equations still continues
199: \cite{CH}.
200: 
201: Inflationary cosmology seems to be less affected by the complex aspects of
202: gravity beyond the semi-classical realm. For inflation to start, a
203: homogeneous patch larger than the Hubble radius, $H^{-1}$, is needed which
204: may require some fine-tuning, especially for inflation at low energy 
205: scales. However, once inflation has begun, the no-hair conjecture
206: \cite{nohair}  should ensure that pre-existing inhomogeneities are driven
207: to  zero. The quantum metric perturbations generated during inflation can 
208: then
209: be tuned to have small  amplitudes $\sim 10^{-5}$ and unless one wants to
210: compute 4-point correlation functions which include graviton loops,
211: comparison with  the anisotropies in the Cosmic Microwave  Background
212: (CMB) is  straight-forward. 
213: 
214: However, nonperturbative production during the coherent oscillation
215: of the inflaton fields~\cite{TB,KLS,Boy,KT,PR}, dubbed
216: preheating~\cite{KLS}, can alter the simple picture of small amplitude 
217: metric  perturbations. Preheating therefore provides a particularly 
218: useful arena for examining some aspects of the boundary 
219: between semi-classical and quantum gravity.
220: 
221: Preheating can lead to the growth of metric perturbations, which
222: in turns is known to stimulate scalar particle creation
223: \cite{TN,massiveB,selfPE} and the production of seed magnetic fields
224: \cite{sting,FG,maroto,BPTV,DPTD}. The amplification of metric 
225: perturbations may also yield
226: runaway instabilities due to the negative specific heat of gravity, 
227: which leads to an interesting possibility of primordial black hole (PBH) 
228: formation \cite{PBHpre1,PBHpre2}. It can also amplify metric
229: perturbations on very large scales
230: \cite{sting,selfBV,selfFB,selfTBV,selfZBS,selfFK}. The growth of metric
231: perturbations, in the case where a PBH does not form, is limited by
232: backreaction effects which are explicitly nonlinear.
233: 
234: Hence on certain scales the Weyl tensor can actually be quite large. This
235: means that there can be significant particle production even if the field
236: is governed by a conformally invariant equation of motion, the classic
237: examples being photons, massless fermions and conformally coupled scalars.
238: In particular, if the effective mass of gravitational relics such as 
239: modulini and gravitini is small enough during reheating, then their  main
240: source of gravitational production will come from metric perturbations. 
241: Since stringent upper bounds hold for the abundances of  these species, it
242: is important to verify that creation from inhomogeneities  (if
243: sufficiently amplified at preheating) does not overcome this threshold.  
244: Our analysis suggests that this does not occur in the simplest models of 
245: chaotic inflation, at least when rescattering is neglected.
246: 
247: Particle creation via inhomogeneities may also be relevant for 
248: alternative scenarios such as the pre-big-bang model \cite{pbb}, where
249: homogeneous gravitational production of moduli and gravitinos is known to
250: be an issue \cite{pbb2} and the Universe goes through a high-curvature
251: phase where higher-order $\alpha'$ corrections are important.
252: 
253: The paper is organized as follows. In section II we review the formalism
254: for computing the number of produced fermions, while in section III we
255: discuss the evolution of metric perturbations during preheating in some
256: specific potentials. Sections IV and V present our numerical results and
257: conclusions respectively.
258: 
259: %
260: \section{Formalism for production of fermions} \label{forma}
261: %
262: 
263: In this section we review the formalism for the production of (Dirac)
264: fermionic quanta by cosmological inhomogeneities, summarizing the
265: calculations of~\cite{Frieman,campos,CV}. 
266: We consider  perturbations around the a FLRW background
267: %
268: \begin{equation}
269: g_{\mu\,\nu}=a^2\left(\eta\right)\,\left(\eta_{\mu\,\nu}
270: +h_{\mu\,\nu}\right)\,\,,
271: \label{gmunu}
272: \end{equation}
273: %
274: where $a(\eta)$ is the scale factor and $\eta$ the conformal time.
275: We will also use proper time, $t$, related to conformal time by
276: $d\eta= dt/a$ and choose to consider only scalar metric 
277: perturbations in the longitudinal
278: gauge~\cite{cosmo,pert} \footnote{Vector perturbations will induce 
279: magnetic fields via the induction equation \cite{TM} but we do not consider
280: them since inflation drives vector perturbations to zero. We also neglect
281: tensor (gravitational wave) modes since they are not strongly amplified 
282: at preheating.}.
283: 
284: At linear order minimally coupled scalar fields 
285: do not induce an anisotropic stress \cite{massiveB} 
286: and hence the metric perturbations 
287: are characterized by a single potential:
288: $h_{\mu\,\nu}=2\,\Phi\,\delta_{\mu\,\nu}\,$, {\em viz}:
289: %
290: \beqn
291: ds^2=a^2(1+2\Phi)d\eta^2
292: -a^2(1-2\Phi)\delta_{ij} dx^i dx^j\,\,.
293: \label{metric}
294: \eeqn
295: %
296: The action of a fermionic field $\psi$ with mass $m$ in 
297: the background~(\ref{metric})  is, to first order in $\Phi$,
298: %
299: \begin{eqnarray}
300: {\cal S} &=& \int\,d^4x\,a^3\,\left(1-2\,\Phi\right)\,\bar{\psi}\,\left\{ 
301: i \,\left[ \, \left( 1- \Phi \right)\, \gamma^0\, \partial_\eta + 
302: \frac{3}{2a} \frac{da}{d\eta}\,\left( 1- \Phi \right)\, 
303: \gamma^0-\frac{3}{2}\frac{d\Phi}{d\eta}\,\gamma^0\right. \right. 
304: \nonumber \\
305: &&\left.\left.+\left( 1+\Phi\right)\, \gamma^i\,\partial_i - \frac{3}{2} 
306: \,\gamma^i\,\partial_i\,\Phi \right]-a\,m\, \right\}\,\psi  \nonumber\\
307: &=& \int d^4x \: {\bar {\tilde{\psi}}} \left[  \left( i \gamma^0 \, 
308: \partial_\eta 
309: + i \gamma^i \, \partial_i - a \, m\right) - 3 \, \Phi \, \gamma^0 \, 
310: \partial_\eta - \Phi \, \gamma^i \, \partial_i - 
311: \frac{3}{2}\frac{d\Phi}{d\eta}\, 
312: \gamma^0 - \frac{3}{2} \, \gamma^i \, \partial_i \Phi \right] \tilde{\psi} 
313:  \nonumber\\
314: &\equiv& 
315: \int d^4x \, \left[ {\cal L}_0 + {\cal L}_I \left( \Phi \right) 
316: \right]\,,
317: \label{feract}
318: \end{eqnarray}
319: %
320: where we have defined $\tilde{\psi} \equiv a^{3/2} \, \psi\,$ and 
321: ${\cal L}_0$, ${\cal L}_I$ denote the homogeneous and interaction 
322: Lagrangians.  
323: 
324: The number density of produced particles is easily computed in the
325: interaction picture, where the evolution of the operators is just
326: determined by the homogeneous expansion of the Universe, and the states 
327: evolve according to the small inhomogeneities in the metric. The field 
328: $\tilde{\psi}$ can be decomposed in the standard way
329: %
330: \begin{equation} \label{decomp0}
331: \tilde{\psi} \left( x \right) = \int \frac{d^3 {\bf k}}{(2\pi )^{3/2}}
332: e^{i {\bf k} \cdot \mathbf{x}}\, \sum_r \left[ u_r \left( k, \eta \right) 
333: \, a_r \left( k \right) + v_r \left( k,\eta \right) b_r^\dagger \left( - k 
334: \right)\right] \,,
335: \end{equation}
336: %
337: where, as usual, the anti-commutation relations are
338: \begin{equation}
339: \left\{a_r(k), a_s^{\dagger}(k') \right\}=
340: \left\{b_r(k), b_s^{\dagger}(k') \right\}=
341: \delta^{(3)} \left( {\bf k}-{\bf k'} \right) \delta_{rs}.
342: \label{anti}
343: \end{equation}
344: %
345: The Fock space is built at the initial time $\eta_i$ starting from the 
346: vacuum state defined by
347: %
348: \begin{equation}
349: a_r \left( k \right) |0 \rangle = b_r \left( k \right) |0 \rangle = 0 \,\,.
350: \label{vacuum}
351: \end{equation}
352: 
353: For massless fermions (the conformally coupled case), the expansion of the 
354: Universe factors out of the free action $\int d^4 x \, {\cal L}_0$ 
355: and the spinors $u_r$ and $v_r$ evolve as in Minkowski spacetime. As a 
356: consequence, 
357: the operators ${a_r}^{(\dagger)} \,,\,{b_r}^{(\dagger)}$ are the physical 
358: annihilation (creation) operators at all times.  Introducing a mass term 
359: breaks the conformal symmetry, and the physical annihilation/creation 
360: operators are defined through the Bogolyubov transformations 
361: (see~\cite{prehferm} for details) %
362: \begin{eqnarray}
363: {\hat a}_r \left(k, \eta \right) &\equiv& \alpha_r \left(k, \eta 
364: \right) \, a_r \left( k \right) - {\beta_r}^* \left(k, \eta \right) \, 
365: {b_r}^\dagger \left( -\,k \right) \,\,, \nonumber\\
366: {{\hat b}_r}^\dagger \left(k, \eta \right) &\equiv& \beta_r
367:  \left( k, \eta \right) \, a_r \left( k \right) + {\alpha_r}^* \left(k, 
368: \eta 
369: \right) \, {b_r}^\dagger \left( -\,k \right) \,\,.
370: \end{eqnarray}
371: %
372: The two Bogolyubov coefficients have initial conditions $\alpha \left( 
373: \eta_i \right) = 1 \,,\, \beta \left( \eta_i \right) = 0\,$ and evolve 
374: according to~\cite{nps}
375: \begin{equation}
376: \alpha_r' = - \frac{m \, k \, a'}{2\, \omega^2} \, {\rm e}^{\,2 \, i 
377: \int^\eta \omega \, d  \eta}\,\beta_r \;\;,\;\; \beta_r' = \frac{m \, 
378: k \, a'}{2\, \omega^2} \, {\rm e}^{\,-\,2\,i \int^\eta \omega \, 
379: d \eta} \alpha_r \,\,,
380: \label{bogo}
381: \end{equation}
382: %
383: with $\omega \equiv \sqrt{k^2 + a^2 \, m^2}\,$. These equations preserve
384: the normalization $\vert \alpha_r \vert^2 + \vert \beta_r \vert^2 = 1\,$, which holds in the fermionic case.
385: 
386: In the interaction picture, the evolution of the initial vacuum (zero
387: particle) state $\vert 0 \rangle$ is determined by the interaction
388: Lagrangian ${\cal L}_I (\Phi)\,$. At first order in $\Phi$ we have
389: %
390: \begin{equation}
391: \vert \psi \rangle = \vert 0 \rangle + \frac{1}{2} 
392: \int d^3 k \, d^3 k' \,\vert k, r \,;\, k', s \rangle\,
393: \langle k, r \,;\, k', s \vert S \vert 0 \rangle \,\,,
394: \label{state}
395: \end{equation}
396: %
397: with the $S$-matrix element ($T$ stands for time ordering)
398: %
399: \begin{equation}
400: \langle k, r \,;\, k', s \vert S \vert 0 \rangle \equiv i \,T\,\langle k, 
401: r 
402: \,;\, k', s \vert {\cal L}_I \left( \Phi \right) \vert 0 \rangle \,\,.
403: \end{equation}
404: 
405: Rigorously speaking, the occupation number can be computed only in the
406: asymptotic future, $\eta_f$, with vanishing perturbations $\Phi \left( 
407: \eta_f \right) = 0\,$. Identical results are obtained for particles and
408: antiparticles, so we concentrate on the former. The expectation value
409: of the number operator ${\hat N} \equiv \left( 2 \, \pi \, a
410: \right)^{-\,3} \int d^3 p \, \hat{a}_r^\dagger \left(p, \eta_f
411: \right) \, \hat{a}_r \left(p, \eta_f \right)$ in the state $\vert
412: \psi \rangle$ is given by the sum of three terms, $N_0 + N_1 + N_2\,$,
413: which are of zeroth, first, and second order in $\Phi\,$, 
414: respectively ~\cite{Frieman,campos,CV},
415: %
416: \begin{eqnarray}
417: N_0 &=& \frac{V}{(2\pi a)^3}\int d^3 k \,
418: \langle 0 \vert \, {\hat a}_{r}^{\dagger}(k) \, {\hat a}_r (k) \, 
419: \vert 0\rangle =
420: \frac{V}{(2\pi a)^3} \int d^3 k \, \vert \beta_k \vert^2 \,\,,
421: \label{N0} \\
422: N_1 &=& - \, \frac{1}{(2\pi a)^3} \int d^3 k
423: \: {\rm Re} \, \left[ \alpha_r \left( k \right) \beta_r \left( k \right) 
424: \, \langle k, r \,;\, -\,k, r \vert S \vert 0 \rangle \right], 
425: \label{N1} \\
426: N_2 &=& \frac{1}{4\,(2\pi a)^3} \int d^3 k \, d^3 k' \, \vert 
427: \langle 0 \vert S \vert k, r  \,;\, k', s \rangle \vert^2 
428: \left[ \vert\alpha_r \left( k \right)\vert^2+\vert\beta_s \left( -\,k' 
429: \right)\vert^2+1 \right] \;\;\;.
430: \label{N2}
431: \end{eqnarray}
432: %
433: The zeroth-order term (\ref{N0}) is the well known expression arising
434: from  the homogeneous expansion of the Universe. $V$ denotes the volume of
435: the Universe at late times, when the perturbations can be neglected. The
436: first-order contribution (\ref{N1}) comes from the combined effect of
437: expansion and inhomogeneities. These two terms vanish for massless
438: fermions, which, as  we remarked, are conformally coupled to the FLRW
439: background (more  explicitly,  we notice that $\beta_r \left(k, \eta
440: \right) = 0$ in this case,  as can be seen from eqs.~(\ref{bogo})).
441: 
442: In the massless case, only the last term~(\ref{N2}) contributes to the 
443: particle production. Furthermore, in this limit $N_2$ acquires the 
444: particularly simple form~\cite{campos}
445: %
446: \begin{equation}
447: N_2 = \frac{1}{160 \, \pi \, a^3} \, \int \frac{d^4 p}{\left( 2 \, \pi 
448: \right)^4}\, \theta \left( p^0 \right) \, \theta \left( p^2 \right) \vert 
449: {\tilde C}^{abcd} \left( p \right) \vert^2 \,\,,
450: \end{equation}
451: %
452: in terms of the Fourier transform 
453: %
454: \begin{equation}
455: {\tilde C}^{abcd} \left( p \right) \equiv \int d^4 x \, {\rm e}^{i\,p\,x} 
456: \, C^{abcd} \left( x \right)
457: \end{equation}
458: %
459: of the Weyl tensor $C^{abcd}\,$.
460: 
461: For the metric~(\ref{metric}), one finds (see also~\cite{campos} for 
462: useful intermediate steps)
463: %
464: \begin{equation}
465: N_2 = \frac{1}{15\,\left( 2 \, \pi \right)^2 \, a^3} \, \int_0^\infty d
466: p_0 \int_{\vert {\bf p} \vert < p_0} d^3 {\bf p} \: \vert {\bf p} \vert^4
467: \: \left| \int_{\eta_i}^{\eta_f} d \eta \, {\rm e}^{i\,p_0\,\eta} \,
468: {\tilde \Phi} \, \left( {\bf p}, \eta \right) \right|^2 \,\,,
469: \label{npsi}
470: \end{equation}
471: %
472: where $\tilde{\Phi}$ is defined by
473: %
474: \begin{equation}
475: \Phi \left( \eta \,,\, {\bf x} \right) \equiv \int \frac{d^3 {\bf 
476: p}}{\left(2\,\pi\right)^{3/2}} \, {\tilde \Phi}
477: \left( {\bf p}, \, \eta \right) \, {\rm e}^{i \, {\bf p} \cdot {\bf x}} 
478: \;\;\;.
479: \label{esse}
480: \end{equation}
481: %
482: In these expressions, $\Phi$ should be regarded as a purely classical 
483: perturbation. The normalization as well as the classicality condition 
484: for $\Phi$ will be discussed in the subsequent sections.
485: 
486: %%
487: \section{Evolution of scalar metric perturbations} \label{secpert}
488: %%
489: 
490: %
491: \subsection{Linearized equations and analytic solutions for 
492: metric perturbations}
493: %
494: 
495: We now discuss the evolution of scalar perturbations $\Phi$ during
496: inflation and reheating in simple models of chaotic
497: inflation~\cite{chaos}, denoting the minimally coupled inflaton field by
498: $\phi$, with potential
499: %
500: \beqn
501: V =\frac12 \, m^2 \, \phi^2 \quad\quad {\rm or} \quad \quad V = \frac14 \, 
502: \lambda \, \phi^4\,.
503: \label{potential}
504: \eeqn
505: %
506: In subsection~\ref{twof} we will then discuss a model where two fields
507: are present.
508: 
509: For the potentials~(\ref{potential}), the COBE normalization
510: of the Cosmic Microwave Background Radiation~\cite{cobe} requires the
511: coupling constants be $m \sim 10^{-6} \, M_p$ for a massive inflaton
512: and $\lambda \sim 10^{-13}$ for the quartic case. Decomposing the
513: inflaton as $\phi (t, {\bf x}) \to \phi (t)+\delta \phi(t, {\bf x})$, the
514: background equations are given by
515: %
516: \begin{eqnarray}
517:   &&H^2 \equiv \left(\frac{\dot{a}}{a}\right)^2=
518:    \frac{8\,\pi}{3\,M_p^2}
519:    \left( \frac12 \dot{\phi}^2+ V \right) \,\,,
520: \label{hubble}\\
521:   &&\ddot{\phi}+3H\dot{\phi}+ \frac{d\,V}{d \, \phi} = 0\,\,,
522: \label{phi}
523: \end{eqnarray}
524: %
525: where dots denote derivatives with respect to physical time. In the
526: numerical simulations which we present in the next section, we will
527: implement the backreaction of field fluctuations within the Hartree
528: approximation. This corresponds to adding the contributions
529: %
530: \begin{eqnarray}
531:  \langle \delta\phi^2 \rangle=\frac{1}{2\pi^2} \int
532: k^2|\delta\phi_k|^2 dk,~~~
533: \langle \delta\dot{\phi}^2 \rangle=\frac{1}{2\pi^2} \int
534: k^2|\delta \dot{\phi}_k|^2 dk,~~~
535: \langle (\nabla\delta \phi)^2 \rangle=\frac{1}{2\pi^2} \int
536: k^4|\delta \phi_k|^2 dk,
537: \label{variance}
538: \end{eqnarray}
539: %
540: to Eqs.~(\ref{hubble}) and (\ref{phi}) (see
541: Refs.~\cite{KLS,Boy} for details). Note that this approach neglects
542: mode-mode coupling and rescattering effects \cite{KT}, which are 
543: important around the end of preheating,
544: and the backreaction effect of metric perturbations.
545:  The Fourier transformed, linearized Einstein
546: equations for field and metric perturbations in the
547: longitudinal gauge are \cite{pert}
548: %
549: \begin{eqnarray}
550: \dot{\Phi}_k + H\Phi_k = \frac{4\,\pi}{M_p^2}\, \dot{\phi}\delta \phi_k, 
551: \label{Phi1}
552: \end{eqnarray}
553: % 
554: \begin{eqnarray}
555: 3H\dot{\Phi}_k + \left( \frac{k^2}{a^2}+3H^2-
556: \frac{4\,\pi}{M_p^2} \, \dot{\phi}^2 \right) \Phi_k=
557: \frac{4\,\pi}{M_p^2} \, \left( \dot{\phi} \delta \dot{\phi}_k+
558: \frac{d^2V}{d \phi^2} \, \phi \delta\phi_k \right),
559: \label{Phi2}
560: \end{eqnarray}
561: % This equation is required to lead to (3.7) [by ST].
562: %
563: \begin{eqnarray}
564: \delta\ddot{\phi}_k + 3H\delta\dot{\phi}_k+
565: \left( \frac{k^2}{a^2} + \frac{d^2V}{d \phi^2}
566: \right) \delta\phi_k=
567: 2(\ddot{\phi}+3H\dot{\phi})\Phi_k+
568: 4\dot{\phi}\dot{\Phi}_k.
569: \label{dphi}
570: \end{eqnarray}
571: 
572: The analytic form of the solutions of the above equations are known in
573: both the  limits of $k \to 0$ and $k \to \infty$ \cite{pert}. During
574: inflation, metric perturbations exhibit  adiabatic growth, $\Phi_k \simeq
575: c \dot{H}/H^2$,  after the first Hubble crossing ($k~\lsim~aH$). During
576: reheating, the super-Hubble modes $(k \ll aH)$ are nearly constant in the
577: single field case. In contrast, the solutions for the small-scale  modes
578: $(k \gg aH)$ can be described by  $\Phi_k \simeq \dot{\phi}\left(c_1e^{i
579: k\eta}+c_2e^{-i k\eta}\right)$, which shows adiabatic damping during
580: reheating.
581: 
582: The system of scalar metric fluctuations $\Phi_k$ and  inflaton
583: fluctuations $\delta \phi_k$ can also be described  in terms of the
584: Mukhanov-Sasaki variable $Q_k$ \cite{muksas}, defined by
585: %
586: \begin{equation}
587: Q_k  \equiv \delta \phi_k + \frac{\dot{\phi}}{H} \Phi_k \,\,,
588: \end{equation}
589: %
590: which satisfies the equation
591: %
592: \begin{equation}
593: \ddot{Q}_k + 3 \, H \, \dot{Q}_k + \left[
594: \frac{k^2}{a^2} +\frac{d^2V}{d \phi^2} + 
595:  \, \left( \frac{\dot{H}}{H} 
596: + 3\, H \right)^\bullet \right] Q_k = 0 \,\,.
597: \label{eqq}
598: \end{equation}
599: %
600: The study of this equation is particularly convenient, since, contrary to 
601: the equation of motion for $\Phi_k$ alone, it is nonsingular during the 
602: oscillations of the inflaton field. The gravitational potential $\Phi_k$ 
603: is then related to $Q_k$ by
604: %
605: \begin{equation}
606: \frac{k^2}{a^2} \, \Phi_k = \frac{4\,\pi}{M_p^2} 
607: \frac{\dot{\phi}^2}{H} \, 
608: \left( \frac{H}{\dot{\phi}} \, Q_k \right)^\bullet \,\,.
609: \label{relqfi}
610: \end{equation}
611: %
612: During inflation, modes of cosmological interest initially (at $t =t_0$)
613: satisfy $k \gg a \, H$, and their equation of motion is that of a free
614: field in an expanding Universe. Quantization of this last quantity is thus
615: straightforward. In the initial vacuum state only positive frequency waves
616: are present (see~\cite{pert} for more details), so that we take as initial
617: conditions
618: %
619: \begin{equation}
620: Q_k \left( t_0 \right) = \frac{1}{a \left( t_0 \right) \, k^{1/2}} \, {\rm 
621: e}^{\,i \alpha_0} \;\;\;,\;\;\; \dot{Q}_k \left( t_0 \right) = 
622: \frac{-\,i\,k^{1/2}}{a^2\left( t_0 \right)} \, {\rm e}^{\,i \alpha_0}\,\,,
623: \label{eqqinitial}
624: \end{equation}
625: %
626: where $\alpha_0$ is an arbitrary phase.
627: 
628: As long as the modes remain much smaller than the Hubble scale 
629: they evolve as plane
630: waves. When their physical lengths grow to of order the Hubble radius 
631: ($k \sim aH$), the variation of the frequency of $Q_k$ becomes important.
632: In fact, resonant amplification of some particular modes can occur during
633: the coherent inflaton oscillations at reheating \cite{TN}. This
634: phenomenon is mostly appreciated when the inflaton is nongravitationally
635: coupled to other fields \cite{selfBV} (see
636: subsection~\ref{twof}). However, some resonant amplification occurs also
637: in the  single-field self coupled case with potential $V=\lambda\phi^4/4$  
638: \cite{selfPE}, as we will discuss 
639: in the next subsection.
640: 
641: %
642: \subsection{Evolution of metric perturbations during reheating}
643: %
644: 
645: %
646: \subsubsection{$V=\frac12 m^2\phi^2$}
647: %
648: 
649: Let us first summarize the behavior of metric fluctuations during
650: reheating in the single field massive inflationary scenario.  During
651: reheating, from the time-averaged relation $\langle \dot{\phi}^2
652: \rangle=\langle m^2\phi^2 \rangle$ one finds $\phi \simeq
653: M_p/(\sqrt{3\pi}\,mt) \sin mt\,$. Neglecting in eq.~(\ref{eqq}) the
654: $H^2$ and $\dot{H}$ terms, which decrease as $\sim t^{-2}$ during
655: reheating, the rescaled variable $\tilde{Q}_k = a^{3/2} \, Q_k\,$ evolves
656: according to a Mathieu equation with time-dependent coefficients
657: %
658: \begin{eqnarray}
659: \frac{d^2}{dz^2}\tilde{Q}_k+(A_k-2q\cos 2z)
660: \tilde{Q}_k=0,
661: \label{Qk2}
662: \end{eqnarray}
663: %
664: where $z=mt/2\,$, and
665: %
666: \begin{eqnarray}
667: A_k=1+\frac{k^2}{(ma)^2},~~~~~
668: q=\frac{1}{\sqrt{2}z}.
669: \label{Aq}
670: \end{eqnarray}
671: %
672: The small $k$ modes ($k~\lsim~ma$) lie in the resonance band around
673: $A_k=1$. However, the cosmic expansion makes $q$ smaller than unity
674: already after one inflaton oscillation, and the resonance is not
675: efficient. As a consequence, modes in the long wavelength limit ($k \to
676: 0$) grow as $\tilde{Q}_k \propto a^{3/2}$ \cite{KH,nata,FB}, which makes
677: $Q_k$ and $\Phi_k$ nearly constant during reheating (for the mode
678: $k~\sim~0.1m$ one has  $|k^{3/2}\Phi_k| \sim 10^{-5}$). On the contrary
679: the sub-Hubble modes ($k~\gsim~m)$ decrease  due to adiabatic damping
680: during reheating and are typically smaller than  the modes which already
681: crossed the Hubble scale  by more than one order of magnitude. This
682: behavior is shown for various wavelengths in Fig.~\ref{massivemet}.
683: 
684: %%%%%%%%%%
685: \begin{figure}
686: \begin{center}
687: \singlefig{10cm}{massivemet.eps}
688: \begin{figcaption}{massivemet}{10cm}
689: The evolution of metric perturbations for $\bar{k}\equiv
690: k/m=0.1, 1, 5, 10$ in the model $V=m^2\phi^2/2$ with inflaton mass
691: $m=10^{-6}M_p$. We start integrating with an initial value,
692: $\phi=0.5M_p$. The scale factor is normalized to unity at this time, and
693: $aH$ is always of order $m$ in the time range plotted. Modes with 
694: $k~\lsim~m$ exhibit adiabatic growth, and then (when
695: reheating starts, i.e., at $mt \approx 1.5$ in the plot)
696: $\Phi_k$ becomes nearly constant. Sub-Hubble modes ($k~\gsim~m$)
697: instead decrease during reheating due to the adiabatic damping of the
698: $\dot{\phi}$ term. 
699: \end{figcaption}
700: \end{center}
701: \end{figure}
702: %%%%%%%%%%
703: 
704: 
705: %
706: \subsubsection{$V=\frac14 \lambda\phi^4$}
707: %
708: 
709: Let us next consider the single field massless chaotic 
710: inflation model, with 
711: potential $V(\phi)=\lambda\phi^4/4$. As shown 
712: in~\cite{Boy,Kaiser,GKLS}, the 
713: evolution of $\phi$ is described (on dropping the term $\frac{1}{a}
714: \frac{d^2a}{d\eta^2}$ from the 
715: equation of motion, eq.~(\ref{phi})) by
716: %
717: \begin{equation}
718: \phi \equiv \frac{\varphi}{a} = \frac{\phi_0}{a} \, {\rm cn} \left( x - 
719: x_0 \,,\, \frac{1}{\sqrt{2}} \right) \,\,,
720: \label{masslessev}
721: \end{equation}
722: %
723: where $x \equiv \sqrt{\lambda} \, \varphi_0 \, \eta$ is the dimensionless
724: conformal time, the suffix $0$ indicates the value of the quantities
725: at the beginning of reheating and ${\rm cn}(x-x_0, 1/\sqrt{2})$ 
726: is the elliptic cosine function.
727: 
728: Using the time-averaged relation $\langle \dot{\phi}^2 \rangle
729: =\lambda\phi^4$, one finds $(\dot{H}/{H})^{\cdot}  \propto
730: (\phi^2)^{\cdot} \propto a^{-3}$ and
731: $\dot{H} \propto a^{-4}\,$. Hence, the Hubble terms in 
732: Eq.~({\ref{eqq}) rapidly become negligible. 
733: Using $x$ and rescaling $\tilde{Q}_k = a \, Q_k$, 
734: Eq.~({\ref{eqq}) is reduced to
735: %
736: \begin{eqnarray}
737: \frac{d^2}{dx^2}\tilde{Q}_k+\left[\kappa^2+
738: 3{\rm cn}^2\left(x, \frac{1}{\sqrt{2}}\right)\right]
739: \tilde{Q}_k=0 \,\,,
740: \label{lame}
741: \end{eqnarray}
742: %
743: where we have defined $\kappa^2 \equiv k^2/ \left( \lambda \, \varphi_0^2
744: \right)\,$, and set $x_0 = 0$ for simplicity. Eq.~({\ref{lame}) is a 
745: Lam\'e equation. From the analytical study of~\cite{GKLS} we deduce the
746: presence of one resonance band in the interval
747: %
748: \beqa 
749: \frac{3}{2}<\kappa^2<\sqrt{3} \,\,,
750: \label{single}
751: \eeqa
752: %
753: characterized by an exponential growth $\tilde{Q}_k \propto e^{\mu_k
754: x}\,$. The maximum value of the growth rate, $\mu_{\rm max}\approx
755: 0.03598\,$, occurs at $\kappa^2 \approx 1.615$. This growth dominates 
756: over the dilution from the expansion of the Universe even for the 
757: unrescaled variable $Q_k\,$, and hence also for $\Phi_k\,$. The growth of
758: $\Phi_k$ continues until the backreaction of inflaton fluctuations shuts
759: off the resonance. The maximal value of the metric perturbation  is found
760: numerically to be $\vert k^{3/2} \, \Phi_k \vert \simeq 5 \times
761: 10^{-\,5}\,$. Modes outside the interval~(\ref{single}) do not exhibit
762: nonadiabatic growth, unless rescattering effects are taken into account
763: \cite{KT,selfPE}. 
764: 
765: 
766: %%%%%%%%%%
767: \begin{figure}
768: \begin{center}
769: \singlefig{10cm}{selfmet.eps}
770: \begin{figcaption}{selfmet}{10cm}
771: The evolution of $\Phi_k$ for $\kappa=0.01, 1.25, 10$ in the single field
772: potential $V=\lambda\phi^4/4$ and $\lambda=10^{-13}$. The initial 
773: values of the numerical evolution are $\phi=0.1M_p\,, a=1\,$. We 
774: include the backreaction effect of produced particles via the Hartree 
775: approximation (see the text for details). Notice the enhancement of the 
776: mode $\kappa=1.25$ in the sub-Hubble resonance band~(\ref{single}).
777: \end{figcaption}
778: \end{center}
779: \end{figure}
780: %%%%%%%%%%
781: 
782: %
783: \subsubsection{$V=\frac14 \lambda\phi^4+\frac12 
784: g^2\phi^2\chi^2$} \label{twof}
785: %
786: 
787: We now discuss the situation in which a massless inflaton $\phi$ is
788: nongravitationally coupled to another scalar field $\chi$, with
789: conformally invariant potential $V=\lambda\phi^4/4+g^2\phi^2\chi^2/2$. 
790: This model was studied in detail, in the absence of
791: metric perturbations, in Refs.~\cite{Kaiser,GKLS}, showing the
792: presence of  an infinite number of strong resonance bands for $\chi$ as a
793: function of
794: \begin{equation}
795: R \equiv g^2/\lambda \,\,.
796: \end{equation}
797: %
798: The conformal invariance allows exact Floquet theory to be used to prove
799: analytically that metric fluctuations can exhibit parametric amplification
800: \cite{selfBV}. This makes the model a very useful testing ground for a
801: number of issues \cite{selfFB,selfTBV,selfZBS,selfFK} including the
802: possible formation of primordial black holes in some parameter space
803: regions \cite{PBHpre2}. Here we briefly summarize those of the above
804: results which are relevant for the calculations of the next section.
805: 
806: As before, it is convenient to introduce a Mukhanov-Sasaki variable for
807: the each field, $Q_k^{(i)}\equiv \delta \varphi_k^{(i)} + \left(
808: \dot{\varphi}^{(i)} / H \right)\Phi_k$, $i = \phi,\chi$ which satisfy a
809: simple generalization of eq.~(\ref{eqq}) \cite{selfBV,selfFB}.
810: 
811: Using eq. (\ref{masslessev}) and rescaling all variables as 
812: ${\tilde F} \equiv a \, F\,$, the equation for
813: $\tilde{Q}_k^{\chi}$ reads \cite{selfBV}
814: %
815: \begin{eqnarray}
816: \frac{d^2}{dx^2}\tilde{Q}_k^{\chi}+\left[\kappa^2+
817: \frac{g^2}{\lambda}{\rm cn}^2\left(x, \frac{1}{\sqrt{2}}\right)\right]
818: \tilde{Q}_k^{\chi}=-2\frac{g^2}{\lambda \phi_0} 
819: {\rm cn}^2\left(x, \frac{1}{\sqrt{2}}\right) \tilde{\chi} 
820: \tilde{Q}_k^{\phi}+M_{\phi\chi}\tilde{Q}_k^{\phi}
821: +M_{\chi\chi}\tilde{Q}_k^{\chi} \,\,,
822: \label{lame2}
823: \end{eqnarray}
824: %
825: where
826: %
827: \beqn
828: M_{\varphi_1\varphi_2}=
829: \frac{8\pi}{a^2M_{p}^2} \left[\frac{1}{aH}
830: \left(\tilde{\varphi}_1'-\frac{a'}{a}\tilde{\varphi}_1
831: \right) \left(\tilde{\varphi}_2'-\frac{a'}{a}\tilde{\varphi}_2
832: \right) \right]' \,\,.
833: \label{M}
834: \eeqn
835: Here prime denotes a derivative with respect to $\eta$.
836: In the early stages of preheating, as long as $\chi$ fluctuations are
837: small relative to $\phi\,$, one can neglect the right hand side of
838: eq.~(\ref{lame2}), which then reduces to the generalized Lam\'e equation
839: studied in~\cite{Kaiser,GKLS} (in the following, we will apply the general
840: analytical results for the Lam\'e equation given in~\cite{GKLS}). Of
841: particular cosmological interest are the longwave modes on the scales
842: proben by the current CMB experiments. It can be shown~\cite{GKLS} that
843: the resonance bands extends to very longwave modes ($\kappa \to 0$) only
844: when the ratio $R$ is in the range
845: %
846: \begin{eqnarray}
847: n \, \left( 2n-1 \right) < R < n \, \left( 2n+1 \right) 
848: \,\,,
849: \label{range}
850: \end{eqnarray}
851: %
852: where $n$ is a positive integer. In each of these intervals, the growth
853: rate $\mu_k$ for $\tilde{Q}_k^{\chi}$ can be analytically found as
854: %
855: \begin{eqnarray}
856: \mu_k=\frac{2}{T} {\rm ln}
857: \left(\sqrt{1+e^{-\pi \epsilon^2}}+e^{-\pi \epsilon^2/2}\right),
858: \label{rate}
859: \end{eqnarray}
860: %
861: where $\epsilon \equiv \sqrt{2/R}\,\kappa^2$ and $T \simeq 7.416$ is the
862: period of inflaton oscillations. Then we have maximal growth when $R$ is
863: exactly $2\,n^2$ and $\kappa=0$, with large characteristic exponent
864: $\mu_{\rm max}=(2/T) {\rm ln} (\sqrt{2}+1)\simeq 0.2377\,$.
865: 
866: It is important to recognize that the exponential growth of
867: $\tilde{Q}_k^{\chi}$ just discussed does not necessarily translate into a
868: parametric amplification of the gravitational potential $\Phi_k\,$. In the
869: two-field case, Eq.~(\ref{Phi1}) is modified to
870: %
871: \begin{eqnarray}
872: \dot{\Phi}_k + H\Phi_k = \frac{4\,\pi}{M_p^2}
873: (\dot{\phi}\delta \phi_k+\dot{\chi}\delta \chi_k) \,\,.
874: \label{Phim}
875: \end{eqnarray}
876: %
877: When $\chi$ and $\delta \chi_k$ are vanishingly small relative to  $\phi$
878: and $\delta \phi_k$, the evolution of $\Phi_k$ is similar to  the single
879: field case. In fact, when $R \gg 1$ the homogeneous mode $\chi$ as well as
880: the longwave $\delta\chi_k$ modes ($k \to 0$) are exponentially suppressed
881: during  inflation \cite{selfBV,selfFB,selfTBV,selfZBS} due to the large
882: effective mass $g \, \phi$ relative to the Hubble parameter (see also
883: Ref.~\cite{suppression,liddle,shinji}). This can be explicitly seen by
884: considering the equation of motion for  $\delta \chi_k\,$, which for
885: super-Hubble modes reads
886: %
887: \begin{equation}
888: \ddot{\delta \chi_k} \, + 3 \, H \, \dot{\delta \chi_k} + g^2 \, \phi^2 \, 
889: \delta \chi_k \simeq 0 \,.
890: \label{delchisup}
891: \end{equation}
892: %
893: Hence, during inflation, the amplitudes of the long-waves modes
894: evolve as \cite{selfBV}
895: $|\delta\chi_k| \propto a^{- \left( 3/2-\nu \right)}\,$, with  
896: $\nu = {\rm Re}\left[9/4 - 3 \, R \,M_p^2 / 
897: \left( 2 \, \pi \, \phi^2 \right) \right]^{1/2}$.
898: This confirms that the inflationary suppression of the long-waves modes
899: becomes more  important as $R$ increases. In addition, considering
900: backreaction in the Hartree approximation, one can show~\cite{selfZBS} that
901: for $R~\gsim~8$ the growth of  field perturbations shuts off the resonance
902: before the super-Hubble metric perturbations ($k \to 0$) begin to be
903: amplified. Hence, for large $R\,$, the evolution of the gravitational
904: potential $\Phi_k$ on the scales probed by CMB experiments does not
905: significantly differ from the ``standard'' one described in~\cite{pert}
906: unless rescattering is taken into account. 
907: For these modes, 
908: parametric amplification is expected to
909: be effective in the first resonance band ~(\ref{range}) even when
910: backreaction is considered~\cite{selfBV,selfFB,selfTBV,selfZBS}.
911: 
912: %%%%%%%%%%
913: \begin{figure}
914: \begin{center}
915: \singlefig{10cm}{selfmetmulti.eps}
916: \begin{figcaption}{selfmetmulti}{10cm}
917: The evolution of metric perturbations $\Phi_k$ and field variances 
918: $\langle\delta\chi^2\rangle$ and $\langle\delta\phi^2\rangle$ for 
919: $\kappa=0.1, 3$ in the potential $V=\lambda\phi^4/4+g^2\phi^2\chi^2/2$  
920: with $g^2/\lambda=2$ and
921: $\lambda=10^{-13}$.  We start integrating with initial values
922: $\phi=0.1M_p$  and $\chi=5 \times 10^{-6}M_p$,
923: corresponding to $\chi=10^{-3}M_p$  at $\phi=4M_p$ 
924: (i.e., 55 e-folds before the end of inflation). 
925: The final results are weakly
926: dependent on the choice of the initial $\chi$ as long as
927: $10^{-6}M_p<\chi<M_p$ at $\phi=4M_p$ \cite{selfZBS,PBHpre2}. 
928: \end{figcaption}
929: \end{center}
930: \end{figure}
931: %%%%%%%%%%
932: 
933: In Fig.~\ref{selfmetmulti} we plot the evolution of $\Phi_k$ for the two
934: modes $\kappa=0.1$ and $\kappa=3$ in the $R=2$ case. We find that metric
935: perturbations begin to grow after the $\delta\chi_k$ fluctuations are
936: sufficiently amplified. The growth of $\Phi_k$ ends around $x=90$ where 
937: the backreaction terminates the resonance, after which perturbations reach
938: the plateau region  found in Fig.~\ref{selfmetmulti}. 
939: 
940: While only long-wave modes are necessary when comparing with 
941: CMB measurements, particle production is also induced by modes $\Phi_k$ 
942: with
943: larger momentum. The sub-Hubble modes ($k~\gsim~aH$) at the beginning of
944: reheating are not severely suppressed during inflation, and therefore the
945: modes $\delta \chi_k$ on those scales can be amplified to large values
946: during preheating. However, from $m_\chi = g \, \langle \phi \rangle$ we
947: see that large values of $R=g^2/\lambda$ results in a suppression of
948: the homogeneous mode $\chi$ during inflation. This also suppresses 
949: amplification of $\Phi$ from the rhs of equation~(\ref{Phim}).
950: Indeed, for $R \gg 1$ the dominant effect for the source of metric
951: perturbations is of second order in nonadiabatic (isocurvature) field
952: perturbations \cite{liddle}. To estimate this growth, we consider the
953: curvature perturbation on uniform-density hypersurfaces \cite{BST,adient},
954: denoted by $\zeta\,$. This quantity is related to the gravitational 
955: potential $\Phi$ via \cite{adient}
956: %
957: \begin{eqnarray}
958: -\zeta={\cal R}+\frac{2\rho}{3(\rho+p)}
959: \left(\frac{k}{aH}\right)^2 \Phi,
960: \label{zeta}
961: \end{eqnarray}
962: %
963: where ${\cal R}\equiv \Phi-(H/\dot{H})(\dot{\Phi}+H\Phi)$ is the comoving
964: curvature perturbation, and $p$ and $\rho$ are the pressure and the energy
965: density, respectively. 
966: The variation of $\zeta$ occurs on large scales in the presence of a 
967: nonadiabatic contribution to the pressure perturbation, $\delta p_{\rm 
968: nad}= \dot{p}\left(\delta p/\dot{p}-\delta\rho/\dot{\rho}\right)$.  One can 
969: estimate the second order contribution to $\zeta$, denoted $\zeta_{(2)}$, 
970: as \cite{liddle} 
971: %
972: \begin{eqnarray}
973: \zeta_{(2)}=\frac{3H}{\dot{\rho}} \int
974: \delta p_{\rm nad}Hdt \simeq
975: \frac{1}{\dot{\phi}^2} \int \left(1+\frac{2\lambda\phi^3}
976: {3H\dot{\phi}}\right) g^2\phi^2 \delta\chi^2 H dt,
977: \label{zetanon}
978: \end{eqnarray}
979: %
980: where we used  the time averaged relation,
981: $\delta\dot{\chi}^2 \simeq m_{\chi}^2 \delta\chi^2 \simeq
982: g^2\phi^2\delta\chi^2$.
983: We find that the second order term, $g^2\phi^2 \delta\chi^2$,
984: can induce a variation of $\zeta$, while the first order term
985: including the homogeneous $\chi$ is vanishingly small for $R \gg 1$.
986: {}From eq.~(\ref{zetanon})  we find
987: %
988: \begin{eqnarray}
989: \left|k^{3/2}\zeta_{(2)}(k) \right|
990: \simeq \frac{\sqrt{2}\pi}{\dot{\phi}^2} \int
991:  \left(1+\frac{2\lambda\phi^3}
992: {3H\dot{\phi}}\right) g^2\phi^2 \sqrt{P_{\delta\chi_k^2}} H dt \,,
993: \label{preper2}
994: \end{eqnarray}
995: %
996: where $P_{\delta\chi_k^2}$ is defined by $P_{\delta\chi_k^2} \equiv
997: \frac{k^3}{2\pi^2} \langle |\delta \chi_k^2|^2 \rangle$.
998: Following Ref.~\cite{liddle}, the power spectrum of the $\delta\chi_k$
999: fluctuation is estimated as 
1000: %
1001: \begin{eqnarray}
1002: P_{\delta\chi_k } \equiv
1003: \frac{k^3}{2\pi^2} \langle |\delta \chi_k|^2 \rangle
1004: \simeq 
1005: \frac{H_0}{g\phi_0} \left( \frac{H_0}{2\pi}\right)^2
1006: \left( \frac{\kappa}{\kappa_0}\right)^3 
1007: \frac{e^{2\mu_k x}}{a^2} \,.
1008: \label{Ppre}
1009: \end{eqnarray}
1010: %
1011: Here $\kappa_0 \simeq 0.15$ is the mode corresponding
1012: to the horizon size at $\phi=0.1M_p$.
1013: {}From eq.~(\ref{Ppre}) one finds
1014: %
1015: \begin{eqnarray}
1016: P_{\delta\chi_k^2} \equiv
1017: \frac{k^3}{2\pi^2} \langle |\delta \chi_k^2|^2 \rangle
1018: \simeq
1019: \frac{k^3}{2\pi} \int_0^{k_c} \frac{P_{\delta\chi}
1020: (|{\bf k'}|) P_{\delta\chi}(|{\bf k-k'}|) }{|{\bf k'}|^3
1021: |{\bf k-k'}|^3} d^3 {\bf k'}
1022: \simeq
1023: \left(\frac{H_0}{g\phi_0}\right)^2 \left( \frac{H_0}{2\pi}\right)^4
1024: \left( \frac{\kappa}{\kappa_0}\right)^3 \frac{M(\kappa, x)}{a^4},
1025: \label{Pdeltachi2}
1026: \end{eqnarray}
1027: %
1028: where
1029: %
1030: \begin{eqnarray}
1031: M(\kappa, x) &\equiv& \left(\frac{\sqrt{\lambda}\phi_0}{\kappa_0}
1032: \right)^3 \int_0^{\kappa_c} d\kappa' \int_0^{\pi} d\theta
1033: e^{2(\mu_{\kappa'}+\mu_{\kappa-\kappa'})x} \kappa'^2
1034: \sin \theta \nonumber \\
1035: &\simeq&
1036: 2 \left(\frac{\sqrt{\lambda}\phi_0}{\kappa_0}
1037: \right)^3 \int_0^{\kappa_c} d\kappa' e^{4\mu_{\kappa'}x}
1038: \kappa'^2 +{\cal O} (\kappa^2).
1039: \label{Meq}
1040: \end{eqnarray}
1041: %
1042: Note that we expanded  the term $\mu_{\kappa-\kappa'}$ around 
1043: $\mu_{\kappa'}$ in eq.~(\ref{Meq}). Due the $\kappa^3$ factor in
1044: eq.~(\ref{Pdeltachi2}), the spectrum $P_{\delta\chi_k^2}$ vanishes  in the
1045: large scale limit ($k \to 0$), which implies that also second order
1046: effects in field perturbations do not induce large metric perturbations on
1047: the scales probed by CMB measurements~\cite{liddle}. In contrast, on
1048: sub-Hubble scales at the beginning of preheating ($\kappa \sim \kappa_0$),
1049: $P_{\delta\chi_k^2}$ is nonvanishing. In this case parametric excitation
1050: of the $\chi$ fluctuation can lead to a growth of $\zeta$ and $\Phi$.
1051: 
1052: %%%%%%%%%%
1053: \begin{figure}
1054: \begin{center}
1055: \singlefig{10cm}{second4.eps}
1056: \begin{figcaption}{second}{10cm}
1057: The evolution of the gravitational potential and the
1058: curvature perturbation for  $\kappa=0.1$  in the model of $V=
1059: \lambda\phi^4/4+g^2\phi^2\chi^2/2$   with $g^2/\lambda=5000$ and
1060: $\lambda=10^{-13}$.  Note that we include the second order effect in field
1061: perturbations. We find that metric perturbations are enhanced on
1062: sub-Hubble scales. We also plot the evolution of the field variance, 
1063: $\langle\delta\chi^2\rangle$ and $\langle\delta\phi^2\rangle$, within
1064: the Hartree approximation.
1065: \end{figcaption}
1066: \end{center}
1067: \end{figure}
1068: %%%%%%%%%%
1069: 
1070: As an example, we plot in Fig.~\ref{second} the evolution of 
1071: $|k^{3/2} \Phi_{(2)}|$ and $|k^{3/2}\zeta_{(2)}|$
1072: for $\kappa=0.1$ in the case of $g^2/\lambda=5000$. 
1073: $|k^{3/2}\Phi_{(2)}|$ is evaluated by making use of eqs.~(\ref{zeta}) 
1074: and (\ref{preper2}).  Note that the final field variances are smaller than 
1075: in the $g^2/\lambda=2$ case, due to the stronger backreaction effect (c.f.  
1076: Fig.~\ref{selfmetmulti}).  The growth of nonadiabatic perturbations 
1077: continues until backreaction ends the resonance.  Although the final value 
1078: of $|k^{3/2}\Phi_{(2)}|$ is somewhat smaller than that of 
1079: $|k^{3/2}\Phi_k|$ in the $g^2/\lambda=2$ case, the fermion 
1080: number density (\ref{npsi}) can be larger due to the contribution of higher 
1081: momentum modes, as we will see in the next section.
1082:  
1083:  
1084: %%
1085: \section{Numerical results for the production of fermions}
1086: %%
1087: 
1088: 
1089: %
1090: \subsection{Quantum to classical transition}
1091: %
1092: 
1093: The results of section~\ref{forma} refer to particle production by a
1094: classical external source. On the contrary, the relation (\ref{relqfi}) of
1095: the previous section relates the modes of the decompositions
1096: %
1097: \begin{equation}
1098: \Phi (\eta,\,{\bf x})=\frac{1}{\sqrt{2}}\,\int\,
1099: \frac{d^3{\bf k}}{\left(2\,\pi\right)^{3/2}}\left[\Phi_k
1100: \left(\eta\right)\,e^{i{\bf 
1101: {k \cdot x}}}\,a(k)+\Phi^*_k\left(\eta\right)\,
1102: e^{-i{\bf {k\cdot x}}}\,a^\dagger(k) \right]\,\,,
1103: \label{deffik}
1104: \end{equation}
1105: %
1106: \begin{equation}\label{defqk}
1107: Q^{(i)}(\eta,\,{\bf x})=\frac{1}{\sqrt{2}}\,\int\,\frac{d^3{\bf 
1108: k}}{\left(2\,\pi\right)^{3/2}}\left[Q_k^{(i)}\left(\eta\right)\,
1109: e^{i{\bf {k\cdot x}}}\,a(k)+Q^*_k{}^{(i)}
1110: \left(\eta\right)\,e^{-i{\bf {k \cdot x}}}\,
1111: a^\dagger (k)\right]\,\,,
1112: \end{equation}
1113: %
1114: to the annihilation and creation operators $a(k)$ and $a^\dagger(k)$. As is
1115: well known, metric perturbations arise from quantum fluctuations during
1116: inflation and some of them eventually undergo a transition to
1117: classicality. A consistent treatment of fermion production requires the
1118: integral (\ref{npsi}) to be limited to those modes of $\Phi$ that have
1119: grown enough to become classical.
1120: 
1121: A simple criterion for classicality is given in Ref.~\cite{PS}.
1122: Classicality is expected to be a good approximation when the following
1123: condition is satisfied:
1124: %
1125: \begin{equation}
1126: \Delta_k^{(i)} \equiv \left| Q_k ^{(i)} \, a  \right| 
1127: \left| ( Q_k^{(i)} \, a )' 
1128: \right| \gg 1 \,\, ~~~~~~~~i = \phi,\chi
1129: \label{class}
1130: \end{equation}
1131: %
1132: The quantity $\Delta_k^{(i)}$ is related to the uncertainty in the
1133: determination of $\langle \Delta Q^{(i)} \, \Delta \Pi^{(i)} \rangle\,$,
1134: where $\Pi^{(i)}$ is the conjugate momentum of $Q^{(i)}\,$. In the case of
1135: pure vacuum fluctuations, the equality $\Delta_k ^{(i)} = 1$ holds, in
1136: accordance with the rules of quantum mechanics. Notice that this is the
1137: case for the initial states~(\ref{eqqinitial}). A much bigger
1138: uncertainty~(\ref{class}) is associated to large fluctuations of classical 
1139: nature.
1140: 
1141: %%%%%%%%%%
1142: \begin{figure}
1143: \begin{center}
1144: \singlefig{10cm}{deltak.eps}
1145: \begin{figcaption}{deltak}{10cm}
1146: The quantum to classical transition as signaled by the evolution of  the
1147: quantity $\Delta_k^{\phi}$ during inflation and subsequent reheating stage
1148: in the single field model,  $V=\lambda\phi^4/4$. The mode $k=a_0H_0$
1149: leaves the Hubble scale  at $\phi=1.5M_p$ (i.e., at the initial time
1150: plotted). $\Delta_k^{\phi}$ grows significantly from  unity after Hubble
1151: radius crossing. In this figure the end of inflation corresponds to 
1152: $\tilde{t}\equiv \sqrt{\lambda}\phi_0 t \approx 10$.
1153: \end{figcaption}
1154: \end{center}
1155: \end{figure}
1156: %%%%%%%%%%
1157: 
1158: %%%%%%%%%%
1159: \begin{figure}
1160: \begin{center}
1161: \singlefig{10cm}{deltamulti.eps}
1162: \begin{figcaption}{deltamulti}{10cm}
1163: The spectra of $\Delta_k^{\chi}$ when the variance 
1164: $\langle \delta \chi^2 \rangle$ reaches its maximum value
1165: in the two-field model of preheating, 
1166: $V=\lambda\phi^4/4+g^2\phi^2\chi^2/2$,
1167: for different values of $R=g^2/\lambda$.
1168: The small $k$ modes ($\kappa<0.5$) are
1169: sufficiently enhanced by parametric resonance 
1170: for $R=2n^2$, which makes the perturbations
1171: highly classical. In the case of $R<20$ we find that 
1172: $\Delta_k^{\chi}$ is very close to unity
1173: for the modes, $\kappa~\gsim~3$.
1174: Therefore the cut-off $\kappa_c=3$ is justified in evaluating 
1175: the occupation number of fermions for  $R<20$.
1176: \end{figcaption}
1177: \end{center}
1178: \end{figure}
1179: %%%%%%%%%%
1180: 
1181: 
1182: In Fig.~\ref{deltak} the time evolution of $\Delta_k^{\phi}$ is  shown in
1183: the model $V = \lambda \, \phi^4/4$, for some particular values of $k\,$.
1184: We notice that $\Delta_k^{\phi}$ begins to increase around the region
1185: where a mode leaves the Hubble scale during inflation. This is expected to
1186: occur quite generically, irrespective of the specific form of the inflaton
1187: potential. On the contrary, modes which are always inside the Hubble scale
1188: typically conserve their quantum nature ($\Delta_k^{\phi} \simeq 1\,$),
1189: unless some physical (model dependent) process drives them classical.
1190: Among these processes, nonperturbative particle creation during 
1191: preheating can drive to a classical level the sub-Hubble modes in the
1192: resonance bands discussed in the previous
1193: section~\cite{KT,PR,selfPE,selfFK}. 
1194: 
1195: In Fig.~\ref{deltamulti} we show instead the spectrum $\Delta_k^\chi$ in
1196: the two-field model of preheating (\ref{twof}) for some values of $R$ 
1197: (the largest time plotted corresponds to the moment at which Hartree
1198: approximation  breaks down).  The case $R=3$ reproduces the results of the
1199: one massless  field case upon identification of $\chi$ with the
1200: fluctuations of the field  $\phi\,$.  For this value,
1201: Fig.~\ref{deltamulti} shows the enhancement of  the modes in the resonance
1202: band~(\ref{single}).  In the two-field case,  strong resonance occurs in
1203: the parameter ranges described by  Eq.~(\ref{range}).  In the center of
1204: the resonance bands, $R=2n^2$, small  $k$ modes ($\kappa~\lsim~0.5$) are
1205: efficiently excited, while the modes  with $\kappa~\gsim~3$ are regarded
1206: as quantum fluctuations  ($\Delta_k^{\chi} \simeq 1$) as long as $R$ is
1207: not much larger than unity  (this evolution does not include rescattering
1208: effects which become  important in the nonlinear regime).
1209: 
1210: In the next section we numerically evaluate the formula for the 
1211: occupation numbers reported in Sec.~II. In the calculation we do not
1212: include those modes which are not excited to a classical  level
1213: ($\Delta_k^{(i)} \simeq 1$). For modes which become classical, from the 
1214: two decompositions~(\ref{esse}) and (\ref{deffik}) we get
1215: %
1216: \begin{equation}
1217: {\tilde \Phi} \left( k \right) = \frac{V^{1/2}}{\sqrt{2}\,\left( 2 \, \pi 
1218: \right)^{3/2}} \, \Phi_k \,\,.
1219: \label{relpert}
1220: \end{equation}
1221: 
1222: 
1223: %%
1224: \subsection{Particle production}
1225: %%
1226: 
1227: We can combine eqs.~(\ref{npsi}) and (\ref{relpert}) to evaluate the
1228: number density of produced fermions in terms  of the modes $\Phi_k$
1229: computed in the previous section. Moreover, we assume the metric
1230: perturbations to be statistically isotropic and homogeneous, so that
1231: $\Phi_{\bf k} = \Phi_{\vert {\bf k} \vert}\,$. In this case, the final
1232: particle density is given by
1233: %
1234: \begin{equation}
1235: N_\psi \equiv \frac{N_2}{V} = \frac{1}{15 \, \left( 2 \, \pi \right)^4 \,
1236: a^3} \int_0^\infty d p_0 \int_0^{p_0} d p \, p^6 \left|
1237: \int_{\eta_i}^{\eta_f} d \eta \, {\rm e}^{i p_0 \eta} \, \Phi_p
1238: \left( \eta \right) \right|^2 \,\,.
1239: \label{npsi2}
1240: \end{equation}
1241: %
1242: As discussed in the previous subsection, the integral over $p \equiv \vert
1243: {\bf p} \vert$ must be restricted to the region where the perturbations
1244: are classical. We have seen in section~\ref{secpert} that this typically
1245: occurs for $p$ smaller than a cut-off value $p_{c}$ which approximately
1246: corresponds to modes which never crossed the Hubble scale (or, in the
1247: massless inflaton case, to the highest resonance band excited). In the
1248: numerical computation of $N_2$ we also integrated $p_0$ up to $p_{c}\,$.
1249: Higher frequencies are not expected to give a sizable contribution to
1250: eq.~(\ref{npsi2}), since the rapidly oscillating phase 
1251: ${\rm e}^{ip_0\eta}$ averages to nearly zero the time integral. 
1252: 
1253: In the numerical calculations we take as the upper limit $\eta_f$ of the
1254: time integral the moment in which backreaction effects start to shut off
1255: the resonance. In the single self-coupling case this occurs at $x_f \equiv
1256: \sqrt{\lambda}\varphi(t_{\rm co})\eta_f \simeq 500$, while in the
1257: two-field case the precise value of $\eta_f$ is a function of the ratio
1258: $R\,$. Note that this is just an approximate calculation, since the above
1259: integral is defined rigorously only for the asymptotic future, $\eta_f \to
1260: \infty$, with vanishing perturbations $\Phi_k \to 0\,$. That is, our
1261: numerical calculation is approximately
1262: %
1263: \begin{equation}
1264: n_{\psi} \equiv a^3 N_\psi \simeq \frac{1}{15 \, \left( 2 \, \pi 
1265: \right)^4} \int_0^{p_c} d p_0 \int_0^{p_0} d p \, p^6 \left| 
1266: \int_{\eta_i}^{\eta_f} d \eta \,
1267: {\rm e}^{i p_0 \eta} \, \Phi_p \left( \eta \right) \right|^2 \,\,.
1268: \label{npsi3}
1269: \end{equation}
1270: %
1271: 
1272: Before presenting the numerical results, we briefly consider the
1273: occupation number of massive fermions produced by the  homogeneous FLRW
1274: expansion, eq.~(\ref{N0}). For sufficiently small masses, the final
1275: occupation number is still (approximately) given by~(\ref{npsi3}), since
1276: the  Bogolyubov coefficient $\beta$ is very close to zero in this limit
1277: (see  eq.~(\ref{bogo}) and (\ref{N2})). The term~(\ref{N0}), although
1278: quadratic in $\beta\,$, is not suppressed by the small perturbations.
1279: Thus, by solving the equation 
1280: \begin{equation}
1281: N_0 \left( m_\psi \right) = N_2 \left( m_\psi = 0  \right)
1282: \end{equation} 
1283: we can estimate up to which mass $m_\psi$ the
1284: productions by the inhomogeneities is comparable with that
1285: from the  homogeneous expansion.
1286: 
1287: The calculation of $N_0$ has been performed numerically in~\cite{igor} 
1288: (see also references therein). For low fermionic masses, the gravitational
1289: production from the homogeneous expansion mainly occurs when the Hubble
1290: expansion rate equals $m_\psi\,$. After this time, the comoving fermion
1291: number density is given by~\cite{igor}
1292: %
1293: \begin{equation}
1294: n_\psi^{\rm hom} \equiv a^3 \, N_0 / V \simeq C_\alpha \, m_\psi^3 \left( 
1295: \frac{H_0}{m_\psi} \right)^{3\,\alpha}\,\,.
1296: \label{hom}
1297: \end{equation}
1298: %
1299: In this expression, $H_0$ denotes the value of the Hubble rate at the
1300: initial time $t_0\,$, where the scale factor $a \left( t_0 \right)$ is
1301: normalized to unity. The parameter $\alpha$ is instead the exponent
1302: appearing in the expansion law $a \left( t \right) \propto t^\alpha$ when
1303: $H \left( t \right) = m_\psi\,$ (as is well known, the coherent inflaton 
1304: oscillations give effective matter domination, $\alpha = 2/3\,$, in the 
1305: $V=m^2\phi^2/2$ case, and effective radiation domination,  $\alpha
1306: =1/2\,$, for $V=\lambda\phi^4/4$). For the first coefficient of
1307: eq.~(\ref{hom}), the two values $C_{2/3} \simeq 3 \times 10^{-\,3}$ and
1308: $C_{1/2} \simeq 10^{-\,2}$ have been numerically found~\cite{igor}.
1309: 
1310: %
1311: \subsubsection{$V=\frac12m^2\phi^2$}
1312: %
1313: 
1314: Let us first present the numerical results for the massive inflaton 
1315: case.  The precise value deduced from the integral~(\ref{npsi3}) is sensitive to 
1316: the cut-off $p_c$.  To discriminate which cut-off should be taken, we 
1317: consider the quantity $\Delta_k^{\phi}$.  We found that $\Delta_k^{\phi}$ 
1318: approaches unity around $k=10m$, so we chose $p_{c}=10 \,m \,$.  In this 
1319: case, for the comoving occupation number (normalizing $a=1$ at $\phi=0.5 \, 
1320: M_p$) we have numerically found
1321: %
1322: \begin{equation}
1323: n_\psi \equiv a^3 N_{\psi} \simeq 3 \times 10^{-14}m^3
1324: \label{nmass}
1325: \end{equation}
1326: %
1327: (just to give an example on how the final result is sensitive to $p_c\,$, 
1328: we found $n_{\psi} \simeq 1 \times 10^{-14}m^3$ for $p_{c}=5m$
1329: and $n_{\psi}\simeq 1 \times 10^{-13}m^3$ for $p_{c}=20m$).
1330: 
1331: Equating (\ref{npsi3}) with (\ref{hom}), the mass $m_{\psi}$ below which 
1332: the production from inhomogeneities dominates over the one from the 
1333: homogeneous expansion is estimated as
1334: %
1335: \begin{equation}
1336: m_{\psi}=\frac{n_{\psi}}{C_{2/3}H_0^2} \simeq 10^2~{\rm GeV}.
1337: \label{mcompar}
1338: \end{equation}
1339: % 
1340: 
1341: Of more physical interest is the fermionic abundance deduced 
1342: from~(\ref{nmass}). By assuming an instantaneous inflaton decay into a 
1343: thermal bath characterized by the reheat temperature $T_{\rm rh}$, 
1344: the particle density divided by the entropy density, $s$, is given by
1345: %
1346: \begin{equation}
1347: Y_\psi \left( T_{\rm rh} \right) \equiv \frac{N_\psi}{s}
1348: \left( T_{\rm rh} \right) 
1349: \simeq 10^{-\,29} \, \frac{T_{\rm rh}}{10^9 \, {\rm GeV}} \,.
1350: \label{Yb}
1351: \end{equation}
1352: %
1353: In the case of other light gravitational relics, such as  gravitini or
1354: modulini, the primordial abundance is severely constrained by the
1355: successful predictions of Big Bang nucleosynthesis. For masses of order
1356: the electroweak breaking scale, the limit is about $Y~\lsim~10^{-\,13}$ --
1357: $10^{-\,12}\,$~\cite{km}. In the present case, eq.~(\ref{Yb}) gives a much
1358: smaller abundance.
1359: 
1360:  
1361: %
1362: \subsubsection{$V=\frac14 \lambda \phi^4$}
1363: %
1364: 
1365: In the quartic case, both super-Hubble modes and the modes in the
1366: interval~(\ref{single}) contribute to the particle production,
1367: eq.~(\ref{npsi3}). From simple estimates (for example combining
1368: eq.~(\ref{npsi3}) with the values for the perturbations shown in
1369: Fig.~\ref{selfmet}), one can see that the main contribution to particle
1370: production is given by the resonance band~(\ref{single}).
1371: 
1372: In calculating the total number densities of produced fermions due to
1373: inhomogeneities, we take the cut-off value in momentum space at
1374: $p_c=3$, over which $\Delta_k^{\phi}$ approaches unity. 
1375: Then the final particle density reached at the
1376: end of preheating is numerically found to be 
1377: %
1378: \begin{equation}
1379: n_{\psi} \simeq 1 
1380: \times 10^{-13}\,(\sqrt{\lambda}\phi_0)^3.
1381: \label{nself}
1382: \end{equation}
1383: %
1384: which is larger than in the massive inflaton case.
1385: 
1386: Proceeding as in the previous subsection we find that the production from 
1387: metric perturbations dominates over that from the homogeneous expansion 
1388: as long as the fermions have masses smaller than
1389: %
1390: \begin{equation}
1391: m_{\psi}=\frac{1}{H_0} \left(\frac{n_{\psi}}{C_{1/2}} \right)^{2/3} \simeq 
1392: 10^5~{\rm GeV} \,\,.
1393: \label{mcompar2}
1394: \end{equation}
1395: %
1396: The final abundance is also greater than in the previous case (thus 
1397: confirming that the integral~(\ref{npsi3}) is dominated by the modes in 
1398: the resonance band),
1399: %
1400: \begin{equation}
1401: Y_\psi \left( T_{\rm rh} \right) \simeq 10^{-\,22} \,\,.
1402: \label{ab2}
1403: \end{equation}
1404: %
1405: Nevertheless it is still too small to have any cosmological effect.
1406: 
1407: 
1408: %
1409: \subsubsection{$V=\frac14 \lambda \phi^4+\frac12g^2\phi^2\chi^2$}
1410: %
1411: Since this potential leads to an infinite number of resonance bands it 
1412: is the case of most interest. 
1413: In Fig.~\ref{npsif} our numerical results for fermionic production 
1414: are presented as a function of $R \equiv g^2/\lambda\,$, for $R \leq 
1415: 10\,$. This is obtained by solving the linearized equation (\ref{Phim})
1416: using the Hartree approximation for field perturbations.
1417:  By comparison with the single field case ($g = 0$, here reproduced by the 
1418: choice $R=3$) the two equations~(\ref{mcompar2}) and (\ref{ab2}) are 
1419: generalized to
1420: %
1421: \begin{equation}
1422: m_\psi \simeq 5 \times 10^{13} \: {\tilde n}_\psi^{2/3} \: {\rm GeV} 
1423: \;\;\;,\;\;\; Y_\psi \left( T_{\rm rh} \right) \simeq 10^{-\,9} \: {\tilde 
1424: n}_\psi \,\,,
1425: \end{equation}
1426: %
1427: where ${\tilde n}_\psi \equiv n_\psi / (\sqrt{\lambda} \, \phi_0 )^3\,$.
1428: 
1429: The qualitative behavior of Fig.~\ref{npsif} is readily understood from the
1430: structure of resonance~(\ref{range}). In the range of $R$ plotted, we have
1431: maximal production when $R$ is in the center of the first resonance band,
1432: in which case $n_\psi$ is found to be $\tilde{n}_\psi \simeq 5 \times
1433: 10^{-10}$, which is about $5000$ times larger than in
1434: the single-field case. In the second resonance band, particle production is
1435: slightly smaller than in the $R=2$ case. 
1436: 
1437: %%%%%%%%%%
1438: \begin{figure}
1439: \begin{center}
1440: \singlefig{10cm}{npsif.eps}
1441: \begin{figcaption}{npsif}{10cm}
1442: The comoving number density, $\tilde{n}_{\psi} = n_{\psi}/
1443: (\sqrt{\lambda}\phi_0)^3$, vs the ratio $g^2/\lambda$, for
1444: $g^2/\lambda \le 10$. Note that this is based on the linear calculation 
1445: using
1446: eq.~(\ref{Phim}), which does not include the second order effect in field
1447: perturbations. We take $\chi=10^{-3}M_p$ at $\phi=4M_p$. We see that
1448: particle creation is enhanced in the super-Hubble 
1449: resonance bands~(\ref{range}).
1450: \end{figcaption}
1451: \end{center}
1452: \end{figure}
1453: %%%%%%%%%%
1454: 
1455: As $R$ increases, the inflationary suppression of $\chi$ becomes more and
1456: more relevant, and particle production from first order 
1457: perturbations is consequently reduced. Nevertheless, the second order
1458: effect of field perturbations described in subsection~III B 3 can lead to 
1459: the
1460: excitation of metric perturbations on sub-Hubble scales. Just for
1461: indicative purposes, we report a couple of our numerical results in this
1462: range. For large $R$, the resonance bands cover only modes up
1463: to~\cite{GKLS}
1464: %
1465: \begin{eqnarray}
1466: \kappa~\lsim~\left(\frac{R}{2\pi^2}\right)^{1/4} \,\,.
1467: \label{kappamax}
1468: \end{eqnarray}
1469: %
1470: In the center of the 10-th instability band, $R=200$, one finds resonant
1471: amplification up to $\kappa~\lsim~1.78$, as can be appreciated in
1472: Fig.~\ref{deltamulti}. By implementing the analysis reported in 
1473: subsection~III B 3
1474: we find $\tilde{n}_{\psi} \simeq 6 \times 10^{-11}$.
1475: As $R$ increases, higher momentum modes contribute  to the growth of
1476: metric perturbations. For example, eq.~(\ref{kappamax}) gives
1477: $\kappa~\lsim~4$ in the $R=5000$ case.  As a consequence, this 
1478: enhances fermionic production. The comoving occupation number 
1479: in this case is  found to be $\tilde{n}_{\psi} \simeq 1 \times 10^{-7} 
1480: \,,\,\, Y_\psi \left( T_{\rm rh} \right)\simeq 10^{-16}\,$.
1481: 
1482: 
1483: It is however important to mention that for $R \gg1$ the whole analysis
1484: becomes very delicate. For large coupling $g\,$, the backreaction effect
1485: of produced particles ends resonant amplification of field perturbations
1486: earlier. In addition, it was found in Ref.~\cite{KT} that the final field
1487: variances get smaller if rescattering of the $\delta\chi$ fluctuations is
1488: taken into account in the rigid FLRW spacetime ($\Phi = 0$). On the
1489: contrary, the effect of rescattering can lead to the excitation of
1490: inflaton fluctuations through the amplification of $\delta\chi$
1491: fluctuations. Since (contrary to the field $\chi$) the homogeneous
1492: inflaton component is not suppressed during inflation, the gravitational
1493: potential can then acquire a potentially large additional source (see
1494: eq.~(\ref{Phim})). Finally, the anisotropic stress is expected to be
1495: important in the nonlinear regime, in which case the relation $\Phi=\Psi$
1496: no longer holds \cite{selfFK}. The study of these effects is certainly
1497: worth separate detailed investigation.
1498: 
1499: 
1500: %
1501: \section{Conclusions}
1502: %
1503: 
1504: 
1505: In the present work we have discussed gravitational creation of fermions by
1506: inhomogeneous scalar perturbations of a Friedmann-Lemaitre-Robertson-Walker
1507: (FLRW) Universe. Massless fermions are conformally coupled to this
1508: background, and thus are not generated by the homogeneous gravitational 
1509: field. As a consequence, gravitational production of light 
1510: particles is solely induced
1511: by metric perturbations which break the conformal flatness of the 
1512: background. 
1513: Although the small size of perturbations on very large scales
1514: indicated by CMB experiments seems to suggest a small production, a precise
1515: computation is still worth while for light gravitational relics.
1516: 
1517: This is particularly true in scenarios in which
1518: metric perturbations undergo parametric amplification during preheating at 
1519: the end of inflation. To
1520: examine this issue more quantitatively, we have studied some particularly
1521: simple models of chaotic inflation. The simplest example is a 
1522: single massive inflaton field ($V = m^2 \, \phi^2
1523: /2$), where parametric resonance is actually absent. In this model, metric
1524: perturbations are nearly constant during reheating for modes which
1525: left the Hubble scale during inflation, whereas the modes deep inside the
1526: Hubble scale exhibit adiabatic damping. As may be expected, particle 
1527: production
1528: is negligibly small in this situation. 
1529: 
1530: We then studied the massless self-coupled inflaton field 
1531: ($V = \lambda \, \phi^4 / 4$) case, in which
1532: parametric resonance occurs for modes in a small band in momentum space 
1533: near 
1534: the Hubble scale.
1535: As a consequence, particle production is enhanced. Moreover, for a
1536: massless inflaton the background energy density decreases more quickly
1537: with respect to the massive case, and this results in a larger abundance
1538: at reheating of the particles generated by the inhomogeneities. Despite 
1539: these two effects, the final result for the production is however still
1540: well below the limit from nucleosynthesis.
1541: 
1542: More efficient production is expected when more fields are present, such 
1543: as the model characterized by the potential
1544: $V=\lambda \, \phi^4/4 + g^2\phi^2 \, \chi^2/2\,$. In this case, the
1545: crucial parameter which determines the resonance bands is $R
1546: \equiv g^2 / \lambda\,$. In particular, the longwavelength modes of the 
1547: second
1548: field $\chi$ are parametrically amplified whenever $R \simeq 2 \, n^2$
1549: with integer $n\,$. At linear order in field perturbations, these modes
1550: are coupled to metric perturbations through a term proportional to 
1551: $\dot{\chi} \delta \chi_k\,$. For relatively small $R\,$, metric 
1552: perturbations
1553: are amplified by this term, and particle production is strongly
1554: enhanced. 
1555: 
1556: For $R > 10$  the large $\chi$ effective mass 
1557: ($m_\chi = g \, \langle \phi \rangle \propto R^{1/2}$) causes the $\chi$ 
1558: field to be exponentialy suppressed during inflation.
1559: As a consequence, for large $R$ metric perturbations mainly increase due to
1560: second order field fluctuations~\cite{liddle}. Particle
1561: production may be expected to become more significant at very large $R\,$.
1562: This is indeed suggested by our numerical results, although in the cases
1563: considered the final abundance is still below the nucleosynthesis
1564: bounds. 
1565: 
1566: Our numerical simulations include the backreaction of the fluctuations on
1567: the background evolution in the Hartree approximation. Rescattering
1568: effects -- mode-mode coupling between fields -- are instead neglected.
1569: This may considerably affect the final abundances, particularly in the
1570: large $R$ limit where the mode coupling becomes particularly important. 
1571: Strong rescattering effects imply an earlier end to preheating, which may
1572: decrease the final values of field and metric perturbations. On
1573: the other hand, rescattering leads to amplification of modes outside the
1574: resonance bands. In particular, inflaton field fluctuations can be
1575: significantly amplified. Since (contrary to the field $\chi$) the 
1576: homogeneous 
1577: inflaton component is not suppressed during inflation, metric
1578: fluctuations may be strongly enhanced by the source term proportional
1579: to $\dot{\phi} \delta\phi_k$ (analogous to the $\dot{\chi} \delta 
1580: \chi_k$ term considered above). This  would be reflected in more
1581: particle creation.
1582: 
1583: Stronger production from inhomogeneities may also occur in different
1584: contexts. In the multi-field inflationary scenario, the gravitational
1585: potential can exhibit nonadiabatic growth during inflation \cite{SY,GW},
1586: especially when the second field $\chi$ has a negative nonminimal coupling
1587: to the Ricci scalar~\cite{TY,STY}. In addition, it was shown that the 
1588: growth
1589: of metric perturbations can be present even in the single field case in
1590: the context of non-slow roll inflation with potential gap~\cite{LS}, and
1591: during graceful exit in string cosmology~\cite{KS}. It may be
1592: interesting to study gravitational particle production in these contexts 
1593: along the lines presented in this work.
1594: 
1595: 
1596: 
1597: %
1598: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1599: \section*{ACKNOWLEDGEMENTS}
1600: The authors  thank J\"urgen Baacke and 
1601: Fermin Viniegra for contributions at early stages of this project
1602: and Robert H. Brandenberger for comments on the draft.
1603: ST also thanks Alexei A. Starobinsky and Jun'ichi Yokoyama
1604: for useful discussions. 
1605: ST is thankful for financial support from 
1606: the JSPS (No. 04942). The work of MP is supported by the European 
1607: Commission RTN programmes HPRN-CT-2000-00148 and 00152. MP thanks Lev 
1608: Kofman for some useful discussions. He also thanks the Canadian Institute 
1609: for Theoretical Astrophysics of Toronto and the Astrophysics group of 
1610: Fermilab for their friendly hospitality during the early stages of this 
1611: work.
1612: 
1613: 
1614: %%%%%%%%%
1615: % references
1616: %%%%%%%%%
1617: 
1618: \begin{thebibliography}{99}
1619: 
1620: \bibitem{BD80}
1621: N. D. Birrell and P. C. W. Davies, {\it Quantum
1622: fields in Curved Space} (Cambridge University Press,
1623: Cambridge, England, 1980).
1624: 
1625: \bibitem{Parker}
1626: L. Parker, Phys. Rev. Lett. {\bf 21}, 562 (1968);
1627: Phys. Rev. {\bf 183}, 1057 (1969).
1628: 
1629: \bibitem{ZS}
1630: Ya. B. Zel'dovich and A. A. Starobinsky, Sov. Phys. 
1631: JETP {\bf 34}, 1159 (1972).
1632: 
1633: \bibitem{HW} 
1634: G. T. Horowitz and R. M. Wald, Phys. Rev. D {\bf 21}, 1462 (1980)
1635: 
1636: \bibitem{Frieman}
1637: J. A. Frieman, Phys. Rev. D {\bf 39}, 389 (1989).
1638: 
1639: \bibitem{campos}
1640: A. Campos and E. Verdaguer, Phys. Rev. D {\bf 45}, 4428 (1992).
1641: 
1642: \bibitem{CV}
1643: J. Cespedes and E. Verdaguer, Phys. Rev. D {\bf 41}, 
1644: 1022 (1990).
1645: 
1646: \bibitem{misner}
1647: C. W. Misner, Phys. Rev. Lett. {\bf 22}, 1071 (1969); 
1648: B. L. Hu and L. Parker, Phys. Rev. D {\bf 17}, 933 (1978).
1649: 
1650: \bibitem{CH}
1651: A. Campos and B. L. Hu, Phys. Rev. D{\bf 58}, 125021 (1998);
1652: A. Campos and E. Verdaguer, Phys. Rev. D{\bf 53}, 1927 (1996).
1653: 
1654: \bibitem{nohair}
1655: G. W. Gibbons and S. W. Hawking, Phs. Rev. D {\bf 15}, 2738 (1977); R. M. 
1656: Wald, Phys. Rev. D{\bf 28}, 2118 (1983);  M. Bruni, F. C. Mena, 
1657: and R. Tavakol,  gr-qc/0107069.
1658: 
1659: \bibitem{TB}
1660: J. Traschen and R. H. Brandenberger, Phys. Rev. D {\bf 42}, 
1661: 2491 (1990); Y. Shatanov, J. Trashen, and R. H. Brandenberger, 
1662: Phys. Rev. D {\bf 51}, 5438 (1995).
1663: 
1664: \bibitem{KLS} 
1665: L. Kofman, A. Linde, and A. A. Starobinsky, 
1666: Phys. Rev. Lett. {\bf 73},  3195 (1994);
1667: Phys. Rev. D {\bf 56}, 3258 (1997).
1668: 
1669: \bibitem{Boy}
1670: D. Boyanovsky et al.,  Phys. Rev. D {\bf 51}, 4419 (1995); 
1671: Phys. Rev. D {\bf 56}, 7570 (1996).
1672: 
1673: \bibitem{KT}
1674: S. Khlebnikov and I. I. Tkachev, Phys. Rev. Lett. {\bf 77}, 219 (1996); 
1675: Phys. Lett. {\bf 390}, 80 (1997);
1676: Phys. Rev. Lett. {\bf 79}, 1607 (1997).
1677: 
1678: \bibitem{PR}
1679: T. Prokopec and T. G. Roos,
1680: Phys. Rev. D  {\bf 55}, 3768 (1997).
1681: 
1682: \bibitem{TN}
1683: A. Taruya and Y. Nambu, 
1684: Phys. Lett. {\bf B428}, 37 (1998).
1685: 
1686: \bibitem{massiveB}
1687: B. A. Bassett, D. I. Kaiser, and R. Maartens, 
1688: Phys. Lett.  {\bf B455}, 84 (1999); 
1689: B. A. Bassett, F. Tamburini, D. I. Kaiser, and
1690: R. Maartens, Nucl. Phys. B {\bf 561}, 188 (1999).
1691: 
1692: \bibitem{selfPE}
1693: R. Easther and M. Parry, 
1694: Phys. Rev. D {\bf 62} 103503 (2000).
1695: 
1696: \bibitem{sting}
1697: B. A. Bassett, C. Gordon, R. Maartens,
1698: and D. I. Kaiser, Phys. Rev. D {\bf 61}, 061302 (R) (2000).
1699: 
1700: \bibitem{FG}
1701: F. Finelli and  A. Gruppuso, 
1702: Phys. Lett. {\bf B502}, 216 (2001). 
1703: 
1704: \bibitem{maroto}
1705: A. L. Maroto, hep-ph/0008288.
1706: 
1707: \bibitem{BPTV}
1708: B. A. Bassett, G. Pollifrone, S. Tsujikawa, and F. Viniegra,
1709: Phys. Rev. D {\bf 63}, 103515  (2001).
1710: 
1711: \bibitem{DPTD}
1712: K. Dimopoulos, T. Prokopec, O. Tornkvist, and A. C. Davis,
1713: astro-ph/0108093.
1714: 
1715: \bibitem{PBHpre1}
1716: A. M. Green and K. A. Malik, 
1717: Phys. Rev. D {\bf 64}, 021301 (2001).
1718: 
1719: \bibitem{PBHpre2}
1720: B. A. Bassett and S. Tsujikawa, 
1721: Phys. Rev. D {\bf 63}, 123503 (2001).
1722: 
1723: \bibitem{selfBV}
1724: B. A. Bassett and F. Viniegra, Phys. Rev. D {\bf 62}, 043507 (2000).
1725: 
1726: \bibitem{selfFB}
1727: F. Finelli and R. H. Brandenberger, Phys. Rev. D {\bf 62}, 083502 (2000).
1728: 
1729: \bibitem{selfTBV}
1730: S. Tsujikawa, B. A. Bassett, and F. Viniegra, JHEP {\bf 08}, 019 (2000).
1731: 
1732: \bibitem{selfZBS}
1733: Z. P. Zibin, R.H. Brandenberger, and D. Scott, 
1734: Phys. Rev. D {\bf 63}, 043511 (2001).
1735: 
1736: \bibitem{selfFK}
1737: F. Finelli and S. Khlebnikov, Phys. Lett. {\bf 504}, 309 (2001).
1738: 
1739: \bibitem{pbb}
1740: M. Gasperini and G. Veneziano, Astropart. Phys. {\bf 1}, 317 (1993).
1741:  
1742: \bibitem{pbb2}
1743: A. Buonanno, M. Lemoine and K. A. Olive, Phys. Rev. D{\bf 62}, 083513 
1744: (2000).
1745: 
1746: \bibitem{cosmo}
1747: J. Bardeen, Phys. Rev. {\bf D22}, 1882 (1980);
1748: H. Kodama and M. Sasaki, Prog. Theor. Phys. Suppl.
1749: No. 78, 1 (1984);.
1750: 
1751: \bibitem{pert} 
1752: V. F. Mukhanov, H. Feldman and R. H. Brandenberger, Phys. Rep. 
1753: {\bf 215}, 203 (1992).
1754: 
1755: \bibitem{TM} 
1756: See e.g., C. G. Tsagas and R. Maartens, Phys. Rev. D {\bf 61}, 
1757: 083519 (2000) and references therein.
1758: 
1759: \bibitem{prehferm} 
1760: G. F. Giudice, M. Peloso, A. Riotto, and I. Tkachev, 
1761: JHEP {\bf 9908}. 014 (1999); 
1762: M. Peloso and L. Sorbo, JHEP {\bf 0005}, 016 (2000);
1763: S. Tsujikawa and H. Yajima, Phys. Rev. D {\bf 64}, 023519 (2001).
1764: 
1765: \bibitem{nps}
1766: H. P. Nilles, M. Peloso and L. Sorbo, JHEP {\bf 0104}, 
1767: 004 (2001).
1768: 
1769: \bibitem{chaos}
1770: A. D. Linde,
1771: Phys. Lett. B {\bf 129}, 177 (1983).
1772: 
1773: \bibitem{cobe} 
1774: G. F. Smoot et al., Ap.J {\bf 396}, L1 (1992); E. L. Wright et 
1775: al., Ap. J. {\bf 396}, L13 (1992).
1776: 
1777: \bibitem{muksas} 
1778: V. F. Mukhanov, JETP Lett. {\bf 41}, 493 (1985);
1779: M. Sasaki, Prog. Theor. Phys. {\bf 76}, 1036 (1986). 
1780: 
1781: \bibitem{KH}
1782: H. Kodama and T. Hamazaki, Prog. Theor. Phys. {\bf 96},
1783: 949 (1996).
1784: 
1785: \bibitem{nata}
1786: Y. Nambu and A. Taruya, 
1787: Prog. Theor. Phys. {\bf 97} 83 (1997).
1788: 
1789: \bibitem{FB}
1790: F. Finelli and R. H. Brandenberger, 
1791: Phys. Rev. Lett. {\bf 82}, 1362 (1999).
1792: 
1793: \bibitem{Kaiser}
1794: D. I. Kaiser, Phys. Rev. D {\bf 53}, 1776 (1995); 
1795: Phys. Rev. D {\bf 56}, 706 (1997); 
1796: Phys. Rev. D {\bf 57}, 702 (1998).
1797: 
1798: \bibitem{GKLS}
1799: P.B. Greene, L. Kofman, A. Linde, and A. A. Starobinsky,
1800: Phys. Rev. D {\bf 56}, 6175 (1997).
1801: 
1802: \bibitem{suppression}
1803: K. Jedamzik and G. Sigl, Phys. Rev. D {\bf 61}, 023519 (2000);
1804: P. Ivanov, Phys. Rev. D {\bf 61}, 023505 (2000).
1805: 
1806: \bibitem{liddle}
1807: A. R. Liddle, D. H. Lyth, K. A. Malik, and D. Wands,
1808: Phys. Rev. D {\bf 61}, 103509 (2000).
1809: 
1810: \bibitem{shinji}
1811: S. Tsujikawa, JHEP {\bf 07}, 024 (2000).
1812: 
1813: \bibitem{BST}
1814: J. M. Bardeen, P. J. Steinhardt, and M. S. Turner,
1815: Phys. Rev. D {\bf 28}, 679 (1983).
1816: 
1817: \bibitem{adient}
1818: C. Gordon, D. Wands, B. A. Bassett, and R. Maartens,
1819: Phys. Rev. D {\bf 63}, 023506 (2001).
1820: 
1821: \bibitem{PS}
1822: D. Polarski and A. A. Starobinsky, Class. Quantum. Grav. 
1823: {\bf 13}, 377 (1996).
1824: 
1825: \bibitem{igor}
1826: V. Kuzmin and I. Tkachev, Phys. Rev. D {\bf 59}, 123006 (1999).
1827: 
1828: \bibitem{km}
1829: M. Kawasaki and T. Moroi, Astrophys.\ J.\  {\bf 452}, 506 (1995).
1830: 
1831: \bibitem{SY} 
1832: A. A. Starobinsky and J. Yokoyama, in: 
1833: Proc. of The Fourth Workshop on General Relativity and Gravitation, 
1834: ed. by K. Nakao {\it et al.} (Kyoto University, 1994), p. 381 
1835: (gr-qc/9502002).  
1836: 
1837: \bibitem{GW}
1838: J. Garc\'{\i}a-Bellido and D. Wands, 
1839: Phys. Rev. D {\bf 52}, 6739 (1995);
1840: D {\bf 53}, 437 (1996).
1841: 
1842: \bibitem{TY}
1843: S. Tsujikawa and H. Yajima, 
1844: Phys. Rev. D {\bf 62}, 123512 (2000); see also
1845: V. Sahni and S. Habib, 
1846: Phys. Rev. Lett. {\bf 81}, 1766 (1998);
1847: B. A. Bassett ans S. Liberati, Phys. Rev. D {\bf 58}, 
1848: 021302 (1998); S. Tsujikawa, K. Maeda, and T. Torii,
1849: Phys. Rev. D {\bf 60}, 063515 (1999); 123505 (1999).
1850: 
1851: \bibitem{STY} 
1852: A. A. Starobinsky, S. Tsujikawa, and J. Yokoyama,
1853: Nucl. Phys. {\bf B610}, 383 (2001). 
1854: 
1855: \bibitem{LS}
1856: A. A. Starobinsky, JETP. Lett. {\bf 55}, 489 (1992);
1857: S. M. Leach, M. Sasaki, D. Wands, and A. R. Liddle, 
1858: Phys. Rev. D {\bf 64}, 023512 (2001).
1859: 
1860: \bibitem{KS}
1861: S. Kawai and J. Soda, 
1862: Phys. Lett. {\bf B 460}. 41 (1999). 
1863: 
1864: \end{thebibliography}
1865: 
1866: \end{document}
1867:  
1868: %%%%%%%%%%%%%%%%%%%%%%
1869: %%%% figures
1870: %%%%%%%%%%%%%%%%%%%%%%
1871: 
1872: