astro-ph0107555/text
1: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2: % Cosmological perturbations from multi-field inflation
3: % in generalized Einstein theories
4: % AS, ST, JY
5: %                 4/10, 7/16 alexei
6: %                 2/18,3/12 jun'ichi
7: %                 2/19 shinji
8: %                 2/27,28,3/2,3/18,4/9, 4/10, 7/29 shinji
9: %
10: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
11: 
12: %\documentstyle[prd,twocolumn,eqsecnum,aps]{revtex}
13: \documentstyle[prd,eqsecnum,aps,epsf]{revtex}
14: %\documentstyle[prd,eqsecnum,aps]{revtex}
15: %\documentstyle[preprint,eqsecnum,aps]{revtex}
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_{\rm pl}}
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: %%%%% singlefig %%%%%
54: \newcommand{\singlefig}[2]{
55: \begin{center}
56: \begin{minipage}{#1}
57: \epsfxsize=#1
58: \epsffile{#2}
59: \end{minipage}
60: \end{center}}
61: %
62: %%%%% figcaption %%%%%
63: \newenvironment{figcaption}[2]{
64:  \vspace{0.3cm}
65:  \refstepcounter{figure}
66:  \label{#1}
67:  \begin{center}
68:  \begin{minipage}{#2}
69:  \begingroup \small FIG. \thefigure: }{
70:  \endgroup
71:  \end{minipage}
72:  \end{center}}
73: %
74: 
75: %------------------------------
76: \def\beq{\begin{equation}}
77: \def\eeq{\end{equation}}
78: \newcommand{\gsim}{\mbox{\raisebox{-1.ex}{$\stackrel
79:      {\textstyle>}{\textstyle\sim}$}}}
80: \newcommand{\lsim}{\mbox{\raisebox{-1.ex}{$\stackrel
81:      {\textstyle<}{\textstyle \sim}$}}}
82: \newcommand{\square}{\kern1pt\vbox{\hrule height
83: 1.2pt\hbox{\vrule width 1.2pt\hskip 3pt
84:    \vbox{\vskip 6pt}\hskip 3pt\vrule width 0.6pt}\hrule
85: height 0.6pt}\kern1pt}
86: %------------------------------
87: 
88: 
89: \begin{document}
90: %\draft \twocolumn[\hsize\textwidth\columnwidth\hsize\csname
91: %@twocolumnfalse\endcsname
92: 
93: \title{Cosmological perturbations from multi-field inflation \\
94: in generalized Einstein theories}
95: \author{Alexei A. Starobinsky$^{1,2}$, Shinji Tsujikawa$^{2,3}$
96: and Jun'ichi Yokoyama$^{4}$}
97: 
98: \address{$^1$ Landau Institute for Theoretical Physics,
99: Russian Academy of Sciences, 117334 Moscow, Russia\\[.3em]}
100: \address{$^2$ Research Center for the Early Universe,
101: University of Tokyo, Hongo, Bunkyo-ku, Tokyo 113-0033, Japan}
102: \address{$^3$ Department of Physics,
103: Waseda University, 3-4-1 Ohkubo, Shinjuku-ku, Tokyo 169-8555,
104: Japan\\[.3em]}
105: \address{$^4$ Department of Earth and Space Science, Graduate School
106: of Science, Osaka University, Toyonaka 560-0043,
107: Japan\\[.3em]}
108:  \date{\today}
109:  \maketitle
110: \begin{abstract}
111: %%ST-2/27-start %%
112: We study cosmological perturbations generated from quantum
113: fluctuations in multi-field inflationary scenarios in generalized
114: Einstein theories, taking both adiabatic and isocurvature modes into
115: account. In the slow-roll approximation, explicit closed-form
116: long-wave solutions for field and metric perturbations are obtained
117: by the analysis in the Einstein frame. Since the evolution of
118: fluctuations depends on specific gravity theories, we make detailed
119: investigations based on analytic and numerical approaches in four
120: generalized Einstein theories: the Jordan-Brans-Dicke (JBD) theory,
121: the Einstein gravity with a non-minimally coupled scalar field, the
122: higher-dimensional Kaluza-Klein theory, and the $R+R^2$ theory with
123: a non-minimally coupled scalar field. We find that solutions obtained
124: in the slow-roll approximation show good agreement with full numerical
125: results except around the end of inflation. Due to the presence of
126: isocurvature perturbations, the gravitational potential $\Phi$ and
127: the curvature perturbations ${\cal R}$ and $\zeta$ do not remain
128: constant on super-horizon scales. In particular, we find that negative
129: non-minimal coupling can lead to strong enhancement of ${\cal R}$ in
130: both the Einstein and higher derivative gravity, in which case it is
131: difficult to unambiguously decompose scalar perturbations into
132: adiabatic and isocurvature modes during the whole stage of inflation.
133: %%ST-2/27-end%%
134: \end{abstract}
135: \vskip 1pc
136: \pacs{pacs: 98.80.Cq}
137: \vskip 2pc
138: %]
139: 
140: %%%%%%%%%%%%%%%%%%%%%%%%%%
141: \section{Introduction}
142: %%%%%%%%%%%%%%%%%%%%%%%%%%
143: 
144: %%ST-2/27-start %%
145: The beauty of the inflationary paradigm is that it both 1) explains
146: why the present-day Universe is approximately homogeneous, isotropic
147: and spatially flat, so that it may be described by the
148: Friedmann-Robertson-Walker (FRW) model in the zero approximation,
149: \cite{inf} and 2) makes detailed quantitative predictions about small
150: deviations from homogeneity and isotropy including density
151: perturbations which produce gravitationally bound objects (such as
152: galaxies, quasars, etc.) and the large-scale structure of the
153: Universe. It is the latter predictions that make it possible to test
154: and falsify this paradigm (for each its concrete realization) like
155: any other scientific hypothesis. Fortunately, all existing and
156: constantly accumulating data, instead of falsifying, continue to
157: confirm these predictions (within observational errors). Historically,
158: among models of inflation making use of a scalar field (called an
159: {\it inflaton}), the original model or the first-order phase
160: transition model \cite{oriinf} failed due to the graceful exit
161: problem, which was taken over by the new \cite{newinf} and the chaotic
162: \cite{chaoinf} inflation scenarios where the inflaton is slowly
163: rolling during the whole de Sitter (inflationary) stage. The latter
164: property was shared by the alternative scenario with higher-derivative
165: quantum gravity corrections \cite{R2} (where the role of an inflaton
166: is played by the Ricci scalar $R$) just from the beginning. Note that
167: a simplified version of this scenario - the $R+R^2$ model - was even
168: shown to be mathematically equivalent to some specific version of the
169: chaotic scenario \cite{W84} (see also a review in \cite{Got92}).
170: 
171: Turning to inhomogeneous perturbations on the FRW background, the
172: inflationary paradigm generically predicts two kinds of them: scalar
173: perturbations and tensor ones (gravitational waves) which are
174: generated from quantum-gravitational fluctuations of the inflaton
175: field and the gravitational field respectively during an inflationary
176: quasi - de Sitter) stage in the early Universe. The spectrum of
177: tensor perturbations generated during inflation was first derived in
178: \cite{St79}, while the correct expression for the spectrum of scalar
179: perturbations {\em after} the end of inflation was obtained in
180: \cite{pert}. For completeness, one should also cite two papers
181: \cite{L80} and \cite{M81} where two important intermediate steps on
182: the way to the right final answer for scalar perturbations were made,
183: in particular, in the latter paper the spectrum of scalar
184: perturbations {\em during} inflation was calculated for the
185: Starobinsky inflationary model \cite{R2}. In order to obtain a small
186: enough amplitude of density perturbations in all the above mentioned
187: slow-roll inflationary models, the inflaton should be extremely
188: weakly coupled to other fields.  It is therefore not easy to find
189: sound motivations to have such a scalar field in particle physics
190: (see, however, \cite{MSYY,LR99}).
191: 
192: Reflecting such a situation, the extended inflation \cite{extinf}
193: scenario was proposed to revive a GUT Higgs field as the inflaton by
194: adopting non-Einstein gravity theories. Although the first version of
195: the inflation model, which considers a first-order phase transition
196: in the Jordan-Brans-Dicke (JBD) theory \cite{BD}, resulted in failure
197: again due to the graceful-exit problem \cite{presc}, it triggered
198: further study of more generic class of inflation models in
199: non-Einstein theories \cite{soft}, in particular extended chaotic
200: inflation \cite{Lin90} where both the inflaton and the Brans-Dicke
201: scalar fields are in the slow rolling regime during inflation. Note
202: that the natural source of Brans-Dicke-like theories of gravity is
203: the low-energy limit of the superstring theory \cite{FT85,C85}
204: with the Brans-Dicke scalar being the dilaton.
205: 
206: Several analyses have been done on the density
207: perturbations produced in extended new or chaotic inflation models
208: \cite{BM,Mc,MM,Der,Garcia}, all of which made use of
209: the constancy of the gauge-invariant quantity $\zeta$ \cite{BST}, or
210: its equivalent ${\cal R}$~\cite{L80,Lyth85} on super-horizon scales,
211: and matched them directly to quantum field fluctuations at the moment
212: of horizon crossing which would be the correct procedure in a
213: single component inflationary model.
214: However, in the presence of two sources of quantum fluctuations
215: (i.e., the inflaton and the Brans-Dicke scalar field),
216: $\zeta$ does not remain constant during inflation due to the
217: appearance of isocurvature perturbations \cite{KS}. In such a case,
218: mixing between adiabatic and isocurvature perturbations
219: may occur due to ambiguity in the definition of the latter ones (see
220: the discussion in the Sec.~III~A below).
221: 
222: Note that it is nothing unusual nor unexpected about non-conservation
223: of the quantities $\zeta$ and ${\cal R}$ even in the long-wave, or
224: super-horizon, $k\ll aH$ limit since they are not conserved integrals
225: of motion even in the one-field, $k\ll aH$ case, if we follow the
226: meaning of this term used in classical mechanics. Here, $a(t)$ is the
227: FRW scale factor, $H\equiv \dot a/a$, where a dot denotes the time
228: derivative, and $k=|\bf k|$ is the conserved covariant momentum of a
229: perturbation Fourier mode. Namely, the ``conservation'' of $\zeta$ in
230: the latter case is restricted to {\em a part} of initial conditions
231: for perturbations for which the decaying mode is not strongly
232: dominating. In the opposite case, since then $\zeta = {\cal O}
233: (k^2\Phi)$, the $k^2$ term in equations for perturbations, e.g. in
234: Eq. (\ref{dotBardeen}) below, may not be neglected even in the
235: $k\ll aH$ case. Of course, any classical dynamical system with $N$
236: degrees of freedom has exactly $2N$ conserved combinations of its
237: coordinates and conjugate momenta irrespective of the fact if it is
238: integrable or not (or even chaotic), but the functional form of these
239: combinations is generically not universal and strongly depends on
240: initial conditions. That is why no special attention is paid to such
241: constant quantities in mechanics. The quantity $\zeta$ is just the
242: examples of such a conserved combination for the growing mode. For
243: the purely decaying mode, another conserved combination may be
244: introduced (see Sec.~III below). It is the specifics of the
245: inflationary scenario where the case of strongly dominated decaying
246: mode is excluded that leads to the impression of the universal
247: conservation of $\zeta$ in the one-field, $k\ll aH$ case. Of
248: course when more modes for each ${\bf k}$ appear in a multi-field
249: case, generic non-conservation of $\zeta$ becomes more transparent.
250: This general remark explains numerous findings of non-conservation of
251: this quantity in particular cases. However, this circumstance does not
252: affect the predictive power of the inflationary paradigm at all since
253: metric perturbations, in particular the gravitational potential
254: $\Phi$, may be calculated during and after inflation without any
255: reference to $\zeta$ conservation or non-conservation. The only
256: problem is that evolution of isocurvature modes of scalar
257: perturbations (in contrast to adiabatic ones) is not universal, and
258: its knowledge requires some additional assumptions about behavior of
259: matter after the end of inflation (in particular, isocurvature modes
260: disappear completely if the total thermodynamic equilibrium is
261: reached at some moment of time).
262: 
263: %%%%%%%%%%%%%%aas%%%%%%%%%%%%%%%
264: 
265: A very important point in the derivation of spectra of primordial
266: perturbations generated in the inflationary scenario is played by
267: exact solutions of perturbations equations in the long-wave, or
268: super-horizon limit $k\to 0$.  It is these solutions that give us a
269: possibility to match quantum inflaton perturbations inside the de
270: Sitter horizon during inflation to scalar perturbations during
271: matter or radiation dominated eras, bypassing the study of
272: physical processes in the Universe in between. That is why such
273: solutions were sought and found for more and more complicated cases
274: in numerous papers. In particular, in \cite{SY}, two of the
275: present authors made the first correct analysis of the issue in the
276: case of slow-roll inflation in the original Brans-Dicke gravity,
277: extending the method used in \cite{St85,PS} to find spectra of all
278: modes of adiabatic and isocurvature fluctuations. This was further
279: used to constrain the parameters of general scalar-tensor theories
280: (e.g., the Damour-Nordvedt model \cite{DN}) from the spectrum of
281: the CMB anisotropy in \cite{GW1,CSY}. General analytic formula
282: for evaluating the spectral index in multi-field inflation was
283: developed in \cite{SS}, which neglected isocurvature modes.
284: $k\to 0$ solutions for all modes in the case of a factorizable
285: potential $U=V_1(\varphi_1)V_2(\varphi_2)$ were found in \cite{GW2}.
286: Finally, a general $k\to 0$ solution for perturbations in the
287: two-field case with an arbitrary potential $U(\varphi_1, \varphi_2)$
288: was found in \cite{multiMS} in the form of a functional over an
289: inflationary background solution, from which explicit solutions for
290: perturbations can be obtained in any case when the background
291: solution may be explicitly integrated in the slow-roll approximation.
292: 
293: However, there exists an old and completely different way to
294: obtain super-horizon ($k\to 0$) solutions for perturbations, even
295: without writing corresponding equations for them:
296: the Lagrange method of variation of the background FRW solution
297: $a(t),\phi_n(t), n=1, ... N$ with respect to all constants
298: entering in it (here $n$ numerates different scalar fields).
299: In the case of matter in the form of classical fields, this
300: method always directly produces all $2N$ physical modes of the
301: $k\to 0$ solution for some quantities (in particular, for some
302: gauge-invariant quantities, too), though to obtain the full
303: $k\to 0$ solution for {\em all} gauge-invariant quantities one has
304: additionally either to use the $0-i$ Einstein equations, or to
305: integrate the $i-j~(i\not= j),i,j=1,2,3$ 
306: Einstein equations.\footnote{For matter in the form of $N$ 
307: hydrodynamic fluids,
308: only $N$ non-decreasing modes, including the most important 
309: growing adiabatic mode, may be obtained in this way.} 
310: The latter necessity
311: is related to the fact that all gauge-invariant quantities
312: constructed from a space-time metric and its first derivatives
313: (including a Newtonian potential $\Phi$) are necessarily non-local,
314: in accordance with the Einstein equivalence principle. As a result,
315: their expressions in terms of quantities which may be defined and
316: measured locally contain $k^2$ in the denominator. So, the next order
317: of the series in powers of $k^2$ (beyond that which follows from the
318: Lagrange method) has to be computed for their determination.
319: 
320: Since the background FRW space-time is in the synchronous form,
321: the Lagrangian method yields $k\to 0$ perturbations in the
322: synchronous gauge, too. Of course, this method produces solutions
323: in the {\em explicit} from only if  the background solution is
324: known explicitly, too. Thus, its power in the two-field case is
325: the same as that of the Mukhanov-Steinhardt functional. On the
326: other hand, it is applicable for any number of scalar fields and
327: during any stage of the Universe evolution, not necessarily at
328: slow-roll inflation.
329: 
330: In the case of adiabatic modes of perturbations (both growing and
331: decaying ones), this method was not only known for many years, but
332: has been already used by Lifshitz and Khalatnikov in a more
333: advanced form -- to produce non-homogeneous solutions near
334: singularity without assuming inhomogeneous perturbations to be
335: small. This is achieved by taking integration constants in the
336: background FRW solution as arbitrary functions of spatial
337: coordinates.  That was how their quasi-isotropic solution~\cite{LH60},
338: or the so called 7-functional solution~\cite{LH63} were constructed.
339: In the case of isocurvature modes of perturbations, this method
340: was recently considered in details in \cite{KH98,ST98,NT98}.
341: 
342: We will use this method in the case of adiabatic modes of
343: perturbations, since it is especially simple in this case. Also, it
344: provides a simple reason for the universal constancy of the
345: properly defined growing adiabatic mode at super-horizon scales
346: (different from that recently proposed in \cite{WMLL}) which does
347: not depend on any local physical process, in particular, on
348: presence or absence of preheating. On the other hand, since the
349: Lagrange method and the Mukhanov-Steinhardt functional are equally
350: powerful in the two-field case in the slow-roll approximation (so
351: that we may use any of them), we will use the latter formula to
352: derive explicit expressions for non-decreasing isocurvature modes of
353: perturbations.
354: 
355: At present there are many generalized Einstein theories which can
356: provide inflationary solutions, e.g, generalized scalar-tensor theory,
357: Einstein gravity with a non-minimally coupled scalar field,
358: higher-dimensional Kaluza-Klein theories, $f(R)$ gravity theories.
359: Making use of the conformal equivalence between these theories,
360: cosmological perturbations can be analyzed in a unified manner (see,
361: e.g., \cite{Hwang}). In the present paper we analyze density
362: perturbations generated in slow-roll inflationary models for a
363: general class of generalized Einstein theories in the presence of two
364: scalar fields. We make use of the conformal transformation~
365: \cite{Maeda89} which transforms the original, or the Jordan frame to
366: the Einstein frame in which equations are somewhat simpler. In
367: particular, the $i-j~(i\not= j)$ Einstein equations directly lead
368: to the universal relation (\ref{equal}) for all models considered
369: in our paper. If this class of gravity theories is considered as a
370: low-energy limit of the superstring theory, then the Jordan frame
371: is also called the string frame.
372: 
373: Though it is always possible to write $k\to 0$ solutions for
374: perturbations in a functional form, actual evolution of perturbations
375: depends on specific gravity theories. In the JBD theory, where the
376: Brans-Dicke parameter is constrained to be $\omega > 3500$ from
377: observations \cite{omega}, the gravitational potential is dominated
378: by adiabatic perturbations \cite{SY,CSY}, in which case the variation
379: of ${\cal R}$ is restricted to be small as we will see later. On the
380: other hand, it was recently found that negative non-minimal coupling
381: with a second scalar field other than inflaton can lead to
382: significant growth of $\zeta$ by the analysis in Jordan frame
383: \cite{TY}. In this case it is not obvious whether the slow-roll
384: analysis provides correct amplitudes of field and metric
385: perturbations, since these exhibit strong enhancement during
386: inflation by negative instability. In this paper we will investigate
387: the validity of slow-roll approximations by numerical simulations in
388: the Einstein frame. We will also work on the evolution of cosmological
389: perturbations in a higher-dimensional theory and the $R+R^2$ theory
390: with a non-minimally coupled scalar field. The latter corresponds to
391: the case where explicit and closed forms of solutions in the
392: super-horizon limit are obtained by slow-roll analysis in spite of
393: a coupled form of the effective potential.
394: 
395: The rest of the present article is organized as follows. In Sec.~II
396: we present the Lagrangian in the Einstein frame and introduce several
397: generalized Einstein theories which can be recasted to the Lagrangian
398: by conformal transformations.  Then in Sec.~III basic equations and
399: closed form solutions for super-horizon perturbations are given. In
400: Sec.~IV we apply the results of Sec.~III to specific gravity theories,
401: namely, the JBD theory, the Einstein gravity with a non-minimally
402: coupled scalar field, the higher-dimensional Kaluza-Klein theory and
403: the $R^2+(1/2)\xi R\chi^2$ theory. In order to confirm analytic
404: estimates, we also show numerical results by solving full equations
405: of motion. We present conclusions and discussions in the final
406: section.
407: 
408: %%%%%%%%%aas%%%%%%%%%%%%%%%%
409: 
410: \section{Inflation in generalized Einstein theories}
411: %%%%%%%%%%%%%%%%%%%%%%%%%%
412: 
413: Consider the following two-field model with scalar fields
414: $\varphi_1$ and $\varphi_2$:
415: %%%%%%%%%%%%%%%
416: \begin{eqnarray}
417: S = \int d^4 x \sqrt{-g} \left[ \frac{1}{2\kappa^2}R- \frac12
418: (\nabla\varphi_1)^2 -\frac12 e^{-2F(\varphi_1)}(\nabla\varphi_2)^2 -
419: U(\varphi_1,\varphi_2) \right],
420: \label{lagrangian}
421: \end{eqnarray}
422: %%%%%%%%%%%%%%%
423: %%ST-3/1-start %%
424: where $\kappa^2/8\pi=G$ is the Newton's gravitational constant,
425: %%ST-3/1-end %%
426: $F(\varphi_1)$ is a function of $\varphi_1$, and
427: $U(\varphi_1,\varphi_2)$ is a potential of scalar fields.
428: Many of the generalized Einstein theories
429: are reduced to the Lagrangian (\ref{lagrangian}) via conformal
430: transformations \cite{Maeda89}.  We have the following theories, which
431: may provide inflationary solutions.
432: %%%%%%%%%%%%%%%
433: \begin{enumerate}
434: \item
435: Theories with a scalar field $\psi$ coupled to gravity whose action
436: is written by
437: %%%%%%%%%%%%%%%
438: \begin{eqnarray}
439: S = \int d^4 x \sqrt{-\hat{g}} \left[ f(\psi)R(\hat{g})- h(\psi)
440: (\hat{\nabla} \psi)^2 -\frac12 (\hat{\nabla}\phi)^2-V(\phi) \right],
441: \label{lag1}
442: \end{eqnarray}
443: %%%%%%%%%%%%%%%
444: where $V(\phi)$ is a potential of inflaton, $\phi$.
445: In this work we consider the following theories.
446: \begin{enumerate}
447: \item Jordan-Brans-Dicke (JBD) theory with a Brans-Dicke field,
448: $\psi$ \cite{BD}.
449: In this case $f$ and $h$ are
450: %%%%%%%%%%%%%%%
451: \begin{eqnarray}
452: f=\frac{\psi}{16\pi},~~~~ h=\frac{\omega}{16\pi \psi},
453: \label{brans}
454: \end{eqnarray}
455: %%%%%%%%%%%%%%%
456: where $\omega$ is the Brans-Dicke parameter which is restricted as
457: $\omega>3500$ from observations \cite{omega}.
458: Making a conformal transformation,
459: %%%%%%%%%%%%%%%
460: \begin{eqnarray}
461: g_{\mu\nu}=\Omega^2\hat{g}_{\mu\nu},
462: \label{conformal}
463: \end{eqnarray}
464: %%%%%%%%%%%%%%%
465: where
466: %%%%%%%%%%%%%%%
467: \begin{eqnarray}
468: \Omega^2=\frac{\kappa^2}{8\pi}\psi \equiv \exp\left(\frac{\kappa
469: \chi}{\sqrt{\omega+3/2}}\right),
470: \label{conformalfac}
471: \end{eqnarray}
472: %%%%%%%%%%%%%%%
473: we obtain the action in the Einstein frame (\ref{lagrangian}) with
474: replacement,
475: %%%%%%%%%%%%%%%
476: \begin{eqnarray}
477: \varphi_1 \to \chi, ~~~\varphi_2 \to \phi,
478: \label{substitute}
479: \end{eqnarray}
480: %%%%%%%%%%%%%%%
481: and
482: %%%%%%%%%%%%%%%
483: \begin{eqnarray}
484: F=(\beta/4)\kappa \chi,~~~~U(\chi,\phi)=e^{-\beta\kappa\chi}V(\phi),
485: \label{brans2}
486: \end{eqnarray}
487: %%%%%%%%%%%%%%%
488: with $\beta=\sqrt{8/(2\omega+3)}$.
489: 
490: \item Non-minimally coupled massless scalar field, $\psi$, with an
491: interaction, $(1/2)\xi R\psi^2$ \cite{TY,SH}.
492: In this case $f$ and $h$ read
493: %%%%%%%%%%%%%%%
494: \begin{eqnarray}
495: f=\frac{1-\xi \kappa^2 \psi^2}{2\kappa^2}, ~~~ h=\frac12.
496: \label{nonmin1}
497: \end{eqnarray}
498: %%jy-3/12 start%%%%%%%%%%%%%
499: Applying the conformal transformation (\ref{conformal}) with
500: $\Omega^2=1-\xi\kappa^2\psi^2$ and
501: defining a new field $\chi$ in order for the kinetic term
502: to be canonical as
503: %%%%%%%%%%%%%%%
504: \begin{eqnarray}
505: \chi=\int \sqrt{\frac{1-(1-6\xi)\xi\kappa^2\psi^2}{(1-\xi\kappa^2
506: \psi^2)^2}}d\psi,
507: \label{nonmin2}
508: \end{eqnarray}
509: %%jy-3/12 end%%%%%%%%%%%%%
510: we obtain the action (\ref{lagrangian}) with replacement
511: (\ref{substitute}) and
512: %%%%%%%%%%%%%%%
513: \begin{eqnarray}
514: F= \frac12 \ln |1-\xi\kappa^2\psi^2|,~~~~
515: U(\chi,\phi)=e^{-4F(\chi)}V(\phi)=\frac{V(\phi)}
516: {(1-\xi\kappa^2\psi^2)^2}.
517: \label{nonmin3}
518: \end{eqnarray}
519: %%%%%%%%%%%%%%%
520: \end{enumerate}
521: The induced gravity theory \cite{induced} is also described by the
522: action, (\ref{lag1}). In this theory the scalar
523: field $\psi$ has its own potential of the form,
524: $V(\psi)=(\lambda/8)(\psi^2-\eta^2)^2$ with $f=(\epsilon/2)\psi^2$
525: and $h=1/2$.
526: 
527: 
528: \item
529: The higher-dimensional theories where the inflaton, $\bar{\phi}$, is
530: introduced in $N=D+4$ dimensions
531: %%%%%%%%%%%%%%%
532: \begin{eqnarray}
533: S = \int d^N x \sqrt{-\bar{g}} \left[\frac{\bar{R}}{2\bar{\kappa}^2} -
534: \frac12(\bar{\nabla}\bar{\phi})^2-\bar{V}(\bar{\phi})\right],
535: \label{higher}
536: \end{eqnarray}
537: %%%%%%%%%%%%%%%
538: where $\bar{\kappa}^2$ and $\bar{R}$ are the $N$-dimensional
539: gravitational constant and a scalar curvature, respectively.
540: We compactify the $N$-dimensional spacetime into the four-dimensional
541: spacetime and the $D$-dimensional internal space with length scale,
542: $b$. Then the metric can be expressed as
543: %%%%%%%%%%%%%%%
544: \begin{eqnarray}
545: ds_N^2=\hat{g}_{\mu\nu}dx^{\mu}dx^{\nu}
546: +b^2 ds^2_D,
547: \label{highermetric}
548: \end{eqnarray}
549: %%%%%%%%%%%%%%%
550: where $\hat{g}_{\mu\nu}$ is a four-dimensional metric. Assuming that
551: extra dimensions are compactified on a torus which has zero curvature,
552: \footnote{Note that there exist other methods of compactifications.
553: One of them is the compactification on the sphere \cite{CW}, in which
554: case stability of extra dimensions and the evolution of
555: cosmological perturbations during inflation are studied in
556: \cite{Amendola,shinji00}.}
557: one gets the following action after dimensional reduction \cite{BM}:
558: %%%%%%%%%%%%%
559: \begin{eqnarray}
560: S=\int d^4 x \sqrt{-\hat{g}} \left(\frac{b}{b_0}\right)^D
561: \frac{1}{2\kappa^2}  \Biggl[
562: \hat{R} +d(d-1)\frac{\partial_{\mu}b \partial_{\nu}b}{b^2}
563: \hat{g}^{\mu\nu}-\bar{\kappa}^2 \left\{ \frac12
564: (\hat{\nabla}\hat{\phi})^2-\hat{V}(\hat{\phi}) \right\} \Biggr],
565: \label{fourdimen}
566: \end{eqnarray}
567: %%%%%%%%%%%%%
568: where $b_0$ is the present value of $b$, and $\hat{R}$ is the scalar
569: curvature with respect to $\hat{g}_{\mu\nu}$.
570: In order to obtain the Einstein-Hilbert action, we make the conformal
571: transformation (\ref{conformal}) with a conformal factor,
572: %%%%%%%%%%%%%%%
573: \begin{eqnarray}
574: \Omega^2=\exp \left(D \frac{\chi}{\chi_0} \right),
575: \label{Omegahigh}
576: \end{eqnarray}
577: %%%%%%%%%%%%%%%
578: where a new scalar field, $\chi$, is defined by
579: %%%%%%%%%%%%%
580: \begin{eqnarray}
581: \chi = \chi_0 {\rm ln} \left( \frac{b}{b_0} \right),~~~~{\rm with}~~~~
582: \chi_0= \left[ \frac{D(D+2)}{2\kappa^2} \right]^{1/2}.
583: \label{chi}
584: \end{eqnarray}
585: %%%%%%%%%%%%%
586: Then the four-dimensional action in the Einstein frame
587: can be described as (\ref{lagrangian}) with replacement
588: (\ref{substitute}) and
589: %%%%%%%%%%%%%%%
590: \begin{eqnarray}
591: F=0,~~~~U(\chi,\phi)=\exp \left(-\beta\kappa \chi \right) V(\phi),
592: ~~~~{\rm with}~~~~\beta=\sqrt{\frac{2D}{D+2}}.
593: \label{FU2}
594: \end{eqnarray}
595: %%%%%%%%%%%%%%%
596: Note that when inflaton is introduced in the four-dimensional action
597: %%jy-3/12 start : instead of (\ref{fourdimen})%%%
598: {\it after} compactification,
599: we find $F=e^{-(\beta/2)\kappa\chi}$ and
600: $U(\chi,\phi)=e^{-\beta \kappa\chi} V(\phi)$ with
601: $\beta=\sqrt{8D/(D+2)}$.
602: %%jy-3/12 end%%%%
603: In this case, however, we do not have inflationary solutions since
604: the effective potential does not satisfy the condition: $\beta<
605: \sqrt{2}$, which is required for power-law inflation to occur (see
606: the next section).
607: 
608: \item
609: The $f(R)$ theories where the Lagrangian includes the higher-order
610: curvature terms, i.e., $\partial f/\partial R$ depends on the scalar
611: curvature $R$ \cite{Hwang}:
612: %%%%%%%%%%%%%%%
613: \begin{eqnarray}
614: S = \int d^4 x \sqrt{-\hat{g}} \left[f(R)-\frac12 (\hat{\nabla}
615: \chi)^2\right].
616: \label{FR}
617: \end{eqnarray}
618: %%%%%%%%%%%%%%%
619: In this case the conformal factor
620: %%%%%%%%%%%%%%%
621: \begin{eqnarray}
622: \Omega^2=2\kappa^2 \left| \frac{\partial f}{\partial R} \right|,
623: \label{dynamical}
624: \end{eqnarray}
625: %%%%%%%%%%%%%%%
626: describes a dynamical freedom in the Einstein-Hilbert action.
627: Introducing a new scalar field
628: %%%%%%%%%%%%%%%
629: \begin{eqnarray}
630: \phi=\sqrt{\frac{3}{2\kappa^2}} {\rm ln}
631: \left[ 2\kappa^2 \left| \frac{\partial f}{\partial R} \right|
632: \right],
633: \label{newphi}
634: \end{eqnarray}
635: %%%%%%%%%%%%%%%
636: the action in the Einstein frame is described as (\ref{lagrangian})
637: with replacement,
638: %%%%%%%%%%%%%%%
639: \begin{eqnarray}
640: \varphi_1 \to \phi, ~~~\varphi_2 \to \chi,
641: \label{substitute2}
642: \end{eqnarray}
643: %%%%%%%%%%%%%%%
644: and
645: %%%%%%%%%%%%%%%
646: \begin{eqnarray}
647: F=\frac{\kappa\phi}{\sqrt{6}},~~~~U(\phi,\chi)= ({\rm sign})
648: \exp\left(-\frac{2\sqrt{6}}{3}\kappa \phi \right) \left[\frac{{\rm
649: (sign)}}{2\kappa^2} R(\phi,\chi) \exp\left(\frac{\sqrt{6}}{3} \kappa
650: \phi\right)-f(\phi,\chi) \right],
651: \label{FU}
652: \end{eqnarray}
653: %%%%%%%%%%%%%%%
654: where ${\rm sign}=(\partial f/\partial R)/|\partial f/\partial R|$.
655: For example, in the $R^2$ theory \cite{R2} with a non-minimally
656: coupled massless
657: $\chi$ field, i.e.,
658: %%%%%%%%%%%%%%%
659: \begin{eqnarray}
660: f(R)=\frac{1}{2\kappa^2}R+\ab R^2-\frac12\xi R\chi^2,
661: \label{R2FR}
662: \end{eqnarray}
663: %%%%%%%%%%%%%%%
664: the effective two-field potential is described as
665: %%%%%%%%%%%%%%%
666: \begin{eqnarray}
667: U(\phi,\chi)=\frac{m_{\rm pl}^4}{(32\pi)^2\ab}
668: e^{-(2\sqrt{6}/3)\kappa\phi}
669: \left( e^{(\sqrt{6}/3)\kappa\phi}-1+\xi\kappa^2\chi^2\right)^2,
670: \label{R2potential}
671: \end{eqnarray}
672: %%%%%%%%%%%%%%%
673: where we have chosen a positive sign.  In this case the $\phi$ field
674: behaves as an inflaton and leads to an inflationary expansion of the
675: Universe.
676: 
677: \end{enumerate}
678: 
679: %%%%%%%%%%aas%%%%%%%%%%%%%%%%
680: 
681: \section{Cosmological perturbations in two-field inflation}
682: %%%%%%%%%%%%%%%%%%%%%%%%%%
683: 
684: \subsection{Basic equations and solution for the adiabatic mode}
685: 
686: Let us first derive the background equations. Variation of the action
687: (\ref{lagrangian}) yields the following background equations for the
688: cosmic expansion rate $H =\dot{a}/a$ and homogeneous parts
689: of scalar fields:
690: %%%%%%%%%%%%%
691: \begin{eqnarray}
692: H^2=\frac{\kappa^2}{3} \left(\frac12 \dot{\varphi}_1^2+\frac12 e^{-2F}
693: \dot{\varphi}_2^2+U \right),
694: \label{hubble1}
695: \end{eqnarray}
696: %%%%%%%%%%%%%
697: %%%%%%%%%%%%%
698: \begin{eqnarray}
699: \dot{H}=-\frac{\kappa^2}{2} \left( \dot{\varphi}_1^2
700: +e^{-2F}\dot{\varphi}_2^2 \right),
701: \label{hubble2}
702: \end{eqnarray}
703: %%%%%%%%%%%%%
704: %%%%%%%%%%%%%
705: \begin{eqnarray}
706: \ddot{\varphi}_1+3H\dot{\varphi}_1+U_{,\varphi_1}
707: +F_{,\varphi_1}  e^{-2F} \dot{\varphi}_2^2=0,
708: \label{varphi1}
709: \end{eqnarray}
710: %%%%%%%%%%%%%
711: %%%%%%%%%%%%%
712: \begin{eqnarray}
713: \ddot{\varphi}_2+3H\dot{\varphi}_2+
714: e^{2F}U_{,\varphi_2}-2F_{,\varphi_1} 
715: \dot{\varphi}_1\dot{\varphi}_2=0,
716: \label{varphi2}
717: \end{eqnarray}
718: %%%%%%%%%%%%%
719: where a prime denotes a derivative with respect to $\varphi_1$.
720: A generic solution of this system contains 5 arbitrary
721: integration constants (two constants appear from the solution of
722: Eq.~(\ref{varphi1}), two -- from Eq.~(\ref{varphi2}), and one
723: from Eq.~(\ref{hubble1}), while Eq.~(\ref{hubble2}) is a consequence
724: of the other equations). However, one of these constants corresponds
725: to a trivial shift of the cosmic time $t$. The Lagrange variation
726: with respect to these constant yields a gauge mode. So, variation
727: with respect to only 4 constants may be used to produce physical
728: solutions for perturbations in the long-wave limit.
729: 
730: Moving to perturbations now, we restrict ourselves to the
731: spatially flat FRW background, first, for simplicity and, second,
732: because recent data on angular fluctuations of the cosmic
733: microwave background (CMB) convincingly confirm the absence of any
734: significant spatial curvature of the Universe within a few percent
735: accuracy (see, e.g., \cite{Boom01}). Then a perturbed space-time
736: metric has the following form for scalar perturbation in an arbitrary
737: gauge:
738: %%%%%%%%%%%%%
739: \begin{eqnarray}
740: ds^2=-(1+2A)dt^2 + 2a(t)B_{,i} dx^idt
741: +a^2(t)[(1+2D)\delta_{ij}+2E_{,i,j}] dx^i dx^j, ~ i,j=1,2,3,
742: \label{metric}
743: \end{eqnarray}
744: %%%%%%%%%%%%%
745: where a comma means usual flat space coordinate derivative and
746: $\Delta$ is the flat $3D$ Laplacian. In the synchronous gauge,
747: $A=B=0,~D=(\lambda + \mu)/6,~\Delta E = -\lambda/2$ in the Lifshitz
748: notations. In the longitudinal gauge, $B=E=0,~A=\Phi,~D=-\Psi$, and
749: $\Phi,\Psi$ are gauge-invariant potentials \cite{Bardeen,MFB,LL93}.
750: Further, we assume the $\exp(i{\bf kx})$ dependence for each
751: Fourier mode ${\bf k}$ and omit the subscript $k\equiv |{\bf k}|$ in
752: expressions for time-dependent parts of perturbations. Note the
753: useful relations between $\lambda,\mu$ and $\Phi,\Psi$, and also
754: between scalar field perturbations in the synchronous gauge
755: $\delta\varphi_{S,n},~n=1,2$ and in the longitudinal gauge
756: $\delta\varphi_n$ (the latter quantities are gauge-invariant
757: actually):
758: \begin{equation}
759: \Phi= -{1\over 2k^2}{d\over dt}(a^2\dot\lambda)~,~~~~
760: \Psi= -{1\over 6}(\lambda+\mu)+{a\dot a\over 2k^2}
761: \dot\lambda~,~~~~
762: \delta\varphi_n = \delta\varphi_{S,n} - {\dot\varphi_n\over 2k^2}
763: \dot\lambda~.
764: \label{relation}
765: \end{equation}
766: Another useful gauge-invariant scalar field perturbation is given
767: by the Mukhanov-Sasaki variable~\cite{MSvar}:
768: \begin{equation}
769: q_n = \delta\varphi_n + {\dot\varphi_n\over H}\Psi =
770: \delta\varphi_{S,n}-{\dot\varphi_n\over 6H}~(\lambda+\mu)~.
771: \label{xi}
772: \end{equation}
773: 
774: For all models considered in our paper, we have the following
775: relation in the Einstein frame
776: \begin{eqnarray}
777: \Phi=\Psi,
778: \label{equal}
779: \end{eqnarray}
780: which follows from the $i-j~(i\not=j)$ Einstein equations taking
781: into account the fact that anisotropic stresses vanish at the linear
782: order there. Transforming back to the Jordan frame, this relation
783: does not hold generically \cite{KS,Hwang}.
784: 
785: Now we can derive one solution for super-horizon perturbations
786: using the Lagrange method without writing equations for
787: perturbations. This is possible since one of integration
788: constants of a background FRW solution, namely, that which appears
789: by integrating Eq. (\ref{hubble1}) is trivial: it is simply a
790: multiplier $a_0$ of $a(t)$. Of course, due to invariance of
791: measurable quantities with respect to equal rescaling of all 3
792: spatial coordinates $x^i$, this constant does not appear in
793: variables like $H(t)$ and $\varphi_n(t)$. So, in the $k\to 0$ limit,
794: \begin{equation}
795: \mu=6{\delta a_0\over a_0}\equiv 3h={\rm const};~~~\lambda,
796: \delta\varphi_{S,n}={\cal O}(k^2h);~~~q_n=
797: -h{\dot\varphi_n\over 2H}.
798: \label{adia}
799: \end{equation}
800: This formula is valid both in the Jordan and the Einstein frames.
801: By definition, this partial solution will be called the {\em
802: growing adiabatic mode} (we will discuss the ambiguity of this
803: definition later). It is clear that this solution exists
804: for {\em any} form of the gravity Lagrangian and the matter
805: energy-momentum tensor. In particular, presence of fast oscillations
806: in a background solution does not affect it, too. The only things
807: which are needed for its existence are the spatial flatness and
808: isotropy of the background metric. That is why the derivation
809: of the spectrum of adiabatic perturbations generated during
810: inflation in the second of Refs.~\cite{pert}, where this method
811: was used even in a more general form (valid when $|h|$ is not
812: necessarily small), does not depend on any physics between
813: inflation and the present era. Note also that the invariance of
814: measurable quantities with respect to different rescaling
815: of spatial coordinates with the total volume fixed leads to
816: the constancy of the non-decreasing (quasi-isotropic) mode of
817: gravitational waves in the super-horizon regime. That property
818: played a crucial role in the derivation of the spectrum of
819: gravitational waves produced during inflation in~\cite{St79}.
820: 
821: Let us now calculate the gauge-invariant quantities $\Phi$ and
822: $\delta\varphi_n$ for the solution (\ref{adia}) in the Einstein
823: frame where the relation (\ref{equal}) holds. In the synchronous
824: gauge, the latter relation reads
825: \begin{equation}
826: \ddot\lambda + 3H\dot\lambda -{k^2\over 3a^2}(\lambda+\mu)=0~.
827: \end{equation}
828: Thus, in the $k\to 0$ limit, we get
829: \begin{equation}
830: \lambda={h\over a^3}\int_{t_1}^ta\,dt~,~~~\Phi=\Psi=-{h\over 2}
831: \left(1-{H\over a}\int_{t_1}^ta\,dt\right)~,~~~
832: \delta\varphi_n=-h{\dot\varphi_n\over 2a}\int_{t_1}^ta\,dt
833: \label{phiadia}
834: \end{equation}
835: ($t_1$ depends on ${\bf k}$ generically). A shift in the $t_1$
836: produces one more superhorizon solution -- the {\em decaying
837: adiabatic mode}:
838: \begin{equation}
839: \lambda=h_1a^{-3},~~~\mu={\cal O}(h_1k^2),~~~\Phi=
840: \Psi=h_1H/2a,~~~
841: \delta\varphi_n=-h_1\dot\varphi_n/2a~.
842: \label{decadia}
843: \end{equation}
844: Thus, the growing adiabatic mode (\ref{adia}) is defined up to
845: an addition of some amount of the decaying adiabatic mode
846: (\ref{decadia}). In the inflationary scenario, $t_1$ is the moment
847: of the first Hubble radius crossing during inflation for each
848: ${\bf k}$, so there is no ambiguity at all. Thus, as a whole
849: the adiabatic solution (mode) is defined unambiguously by Eqs.~
850: (\ref{adia},~\ref{phiadia}). On the other hand, one may always add
851: some amount of the adiabatic mode to other, isocurvature solutions
852: (modes). So, the definition of isocurvature modes is not unique.
853: In particular, this ambiguity may be used to make $\Phi=\Psi=0$
854: for an isocurvature mode in the Einstein frame at some chosen
855: moment of time, e.g., at the end of inflation or at the moment
856: when the full thermal equilibrium is reached (if the latter occurs
857: at all). Of course, this choice does not affect any observable
858: quantities.
859: 
860: A useful gauge-invariant quantity is the comoving curvature
861: perturbation ${\cal R}$~\cite{L80,Lyth85}:
862: \begin{equation}
863: {\cal R}= -{1\over 6}(\lambda+\mu)+{H\over 6\dot H}
864: (\dot\lambda + \dot\mu)=\Psi-\frac{H}{\dot{H}}(\dot\Psi+H\Phi)~.
865: \label{calR}
866: \end{equation}
867: A similar quantity is the curvature perturbation on uniform
868: density hypersurfaces introduced in~\cite{BST}:
869: \begin{equation}
870: \zeta=D-H{\delta\rho\over\dot\rho}=\left({1\over 6}-{k^2\over
871: 18a^2\dot H}\right)(\lambda+\mu)-{H\over 6\dot H}\dot\mu
872: =-{\cal R}+{k^2\over 3a^2\dot H}\Psi
873: \label{Bardeen}
874: \end{equation}
875: where $\rho$ is the total energy density of matter and
876: $\delta\rho$ -- its perturbation. For the solution (\ref{adia}),
877: we get
878: \begin{equation}
879: {\cal R}=-\,\zeta=-h/2={\rm const}\not=0.
880: \label{constant}
881: \end{equation}
882: Thus, we have the theorem: \\
883: {\it In the superhorizon limit} $k\to 0$ {\it and for $\dot H\not=0$
884: identically, there always exists one solution for perturbations (the
885: growing adiabatic mode) for which} ${\cal R}=-\zeta$ {\it is
886: conserved in the leading order in} $k^2$ {\it (apart from the
887: vicinity of points where} $\dot H=0$). \\
888: Let us emphasize that this statement is valid both in the Jordan
889: and the Einstein frames.
890: 
891: Pathological behavior of ${\cal R}$ and $\zeta$ at the moments
892: of time when $\dot H=0$ where they diverge (if terms of the next
893: order in $k^2$ are taken into account) is solely an artifact of
894: their definition. The potentials $\Phi$ and $\Psi$ remain regular
895: and small near these points, so no consideration of non-linear
896: effects is required there. Moreover, it immediately follows from
897: the formulas (\ref{adia},~\ref{phiadia}) that the same constant
898: value (\ref{constant}) of ${\cal R}$ and $\zeta$ is restored
899: after passing through any point where $\dot H=0$. This behavior
900: of $\zeta$ is clearly seen in our numerical calculations in
901: Figs. 1 and 3 below.
902: 
903: What about conservation of $\cal R$ and $\zeta$ for the
904: decaying adiabatic mode (\ref{decadia}) ? Here, there is a subtle
905: point. In the leading order, ${\cal R}=\zeta=0$ for this mode.
906: However, one may not say that these quantities are conserved (even
907: approximately). Real (approximate) conservation would require
908: $H|\dot {\cal R}/{\cal R}|\ll 1$ that is not valid if the decaying
909: adiabatic mode is strongly dominating \cite{Leach}. 
910: Also, the relation ${\cal R}
911: \approx -\zeta$ is not correct for this mode. On the other hand,
912: in the Einstein frame one can introduce the gauge invariant quantity
913: ${\cal T}\equiv a\Phi/H$ which does conserve in the super-horizon
914: limit in the leading order in $k^2$.
915: 
916: Finally, for other solutions (isocurvature modes) ${\cal R}$ and
917: $\zeta$ are not conserved in the super-horizon limit, too. This can
918: be easily seen by considering the time derivative of ${\cal R}$
919: \cite{GW1,GW2}:
920: %%%%%%%%%%%%%
921: \begin{eqnarray}
922: \dot{{\cal R}}=\frac{H}{\dot{H}}\frac{k^2}{a^2}\Phi
923: +H\left(\frac{\delta\varphi_1}{\dot{\varphi_1}}-
924: \frac{\delta\varphi_2}{\dot{\varphi_2}}\right) Z,
925: \label{dotBardeen}
926: \end{eqnarray}
927: %%%%%%%%%%%%%
928: where
929: %%%%%%%%%%%%%
930: \begin{eqnarray}
931: Z \equiv \frac{2e^{-2F}\dot{\varphi}_1\dot{\varphi}_2
932: (\ddot{\varphi}_1\dot{\varphi}_2-\dot{\varphi}_1\ddot{\varphi}_2
933: +\dot{F}\varphi_1\varphi_2)+F_{,\varphi_1}
934: \dot{\varphi}_1e^{-4F}\dot{\varphi}_2^4}{(\dot{\varphi}_1^2+e^{-2F}
935: \dot{\varphi}_2^2)^2}.
936: \label{Z}
937: \end{eqnarray}
938: %%%%%%%%%%%%%
939: The quantity
940: %%%%%%%%%%%%%
941: \begin{eqnarray}
942: S_{\varphi_1\varphi_2} \equiv
943: H\left(\frac{\delta\varphi_1}{\dot{\varphi_1}}-
944: \frac{\delta\varphi_2}{\dot{\varphi_2}}\right),
945: \label{entropy}
946: \end{eqnarray}
947: %%%%%%%%%%%%%
948: represents a generalized entropy perturbation between $\varphi_1$ and
949: $\varphi_2$ fields\cite{GWBM,HN}. In the multi-field case,
950: $S_{\varphi_1\varphi_2} \ne 0$ and $Z \ne 0$ generically. So, the
951: presence of isocurvature perturbations leads to the variation of
952: ${\cal R}\approx - \zeta$.
953: 
954: Note also that we use the terms "adiabatic perturbations" and
955: "isocurvature perturbations", as is commonly done, to denote
956: different {\em solutions} (modes) of the same physical variables.
957: This should be contrasted with the approach of 
958: the recent paper \cite{GWBM} where "adiabatic" 
959: and "entropy" {\em fields} are
960: introduced which are certain linear combinations of initial fields
961: $\varphi_n(t)$, and then adiabatic and entropy perturbations
962: mean perturbations of these fields. However, for our adiabatic
963: solution (\ref{adia},~\ref{phiadia}) the perturbed entropy field
964: $\delta s\propto(\dot\varphi_1\delta\varphi_2-\dot\varphi_2\delta
965: \varphi_1)=0$ in the super-horizon limit, even if the background
966: field trajectory in the scalar field space is curved. Thus, the
967: adiabatic mode is not sourced by isocurvature (entropy) modes, in
968: contrast to results of \cite{GWBM}.
969: 
970: That is all what may be obtained without solving equations for
971: perturbations. So, to proceed further, equations for time-dependent
972: parts of metric and field fluctuations have to be written. They
973: read:
974: %%%%%%%%%aas%%%%%%%%%%%%%%%%
975: \begin{eqnarray}
976: \ddot{\Phi}+4H\dot{\Phi}+\kappa^2U\Phi= \frac{\kappa^2}{2} 
977: \left[ \dot{\varphi}_1\delta\dot{\varphi}_1-
978: (U_{,\varphi_1}+F_{,\varphi_1} 
979: e^{-2F}\dot{\varphi_2}^2) \delta\varphi_1+
980: e^{-2F}\dot{\varphi}_2\delta\dot{\varphi}_2-
981: U_{,\varphi_2}\delta\varphi_2 \right],
982: \label{perturb4}
983: \end{eqnarray}
984: %%%%%%%%%%%%%
985: %%%%%%%%%%%%%
986: \begin{eqnarray}
987: \left(\frac{k^2}{a^2}-\dot{H}\right)\Phi
988: =-\frac{\kappa^2}{2}\left[\dot{\varphi}_1\delta\dot{\varphi}_1
989: +(3H\dot{\varphi}_1+U_{,\varphi_1}-F_{,\varphi_1}e^{-2F}
990: \dot{\varphi}_2^2) \delta\varphi_1+
991: e^{-2F}\dot{\varphi}_2\delta\dot{\varphi}_2+
992: (U_{,\varphi_2}+3H\dot{\varphi}_2e^{-2F}) \delta\varphi_2 \right],
993: \label{perturb5}
994: \end{eqnarray}
995: %%%%%%%%%%%%%
996: %%%%%%%%%%%%%
997: \begin{eqnarray}
998: \dot{\Phi}+H\Phi=\frac{\kappa^2}{2}
999: \left( \dot{\varphi}_1 \delta \varphi_1
1000: +e^{-2F}\dot{\varphi}_2 \delta \varphi_2 \right),
1001: \label{perturb1}
1002: \end{eqnarray}
1003: %%%%%%%%%%%%%
1004: %%%%%%%%%%%%%
1005: \begin{eqnarray}
1006: \delta\ddot{\varphi}_1&+& 3H\delta\dot{\varphi}_1
1007: +\left[\frac{k^2}{a^2}+U_{,\varphi_1\varphi_1}-
1008: \left(e^{-2F}\right)_{,\varphi_1\varphi_1}
1009:  \frac{\dot{\varphi}_2^2}{2} \right] \delta\varphi_1
1010: + 2F_{,\varphi_1} e^{-2F} \dot{\varphi}_2
1011: \delta\dot{\varphi}_2+U_{,\varphi_1\varphi_2}\delta\varphi_2 \nonumber \\
1012: &=& 4\dot{\varphi}_1 \dot{\Phi}-2U_{,\varphi_1}\Phi,
1013: \label{perturb2}
1014: \end{eqnarray}
1015: %%%%%%%%%%%%%
1016: %%%%%%%%%%%%%
1017: \begin{eqnarray}
1018: \delta\ddot{\varphi}_2 &+& (3H-2\dot{F})\delta\dot{\varphi}_2
1019: +\left(\frac{k^2}{a^2}+e^{2F}U_{,\varphi_2\varphi_2} \right)
1020: \delta\varphi_2-2\dot{F}\delta\dot{\varphi}_1+e^{2F}
1021: \left( 2F_{,\varphi_1}
1022: U_{\varphi_2}+U_{\varphi_1 \varphi_2} \right) \delta \varphi_1
1023: \nonumber \\
1024: &=& 4\dot{\varphi}_2 \dot{\Phi}-2e^{2F}U_{,\varphi_2}\Phi.
1025: \label{perturb3}
1026: \end{eqnarray}
1027: %%%%%%%%%%%%%
1028: The relation (\ref{perturb5}) clearly indicates that metric
1029: perturbations are determined when the evolution of scalar fields are
1030: known.
1031: 
1032: \subsection{Closed form solutions in slow-roll approximations}
1033: 
1034: The use of the slow-roll approximation allows us to obtain closed form
1035: solutions for isocurvature perturbations in the long-wave limit
1036: \cite{SY,GW1,GW2,multiMS}.
1037: Under this approximation, the background equations are simplified as
1038: %%%%%%%%%%%%%
1039: \begin{eqnarray}
1040: H^2=\frac{\kappa^2}{3} U,
1041: \label{hubbleslow}
1042: \end{eqnarray}
1043: %%%%%%%%%%%%%
1044: %%%%%%%%%%%%%
1045: \begin{eqnarray}
1046: 3H\dot{\varphi}_1+U_{,\varphi_1}=0,
1047: \label{varphi1slow}
1048: \end{eqnarray}
1049: %%%%%%%%%%%%%
1050: %%%%%%%%%%%%%
1051: \begin{eqnarray}
1052: 3H\dot{\varphi}_2+e^{2F}U_{,\varphi_2}=0.
1053: \label{varphi2slow}
1054: \end{eqnarray}
1055: %%%%%%%%%%%%%
1056: Combining Eqs.~(\ref{hubbleslow})-(\ref{varphi2slow}) with
1057: Eq.~(\ref{hubble2}), we find
1058: %%%%%%%%%%%%%
1059: \begin{eqnarray}
1060: \dot{\varphi}_1=-\frac{H}{\kappa^2}
1061: \frac{U_{,\varphi_1}}{U},~~~~~
1062: \dot{\varphi}_2=-\frac{He^{2F}}{\kappa^2}
1063: \frac{U_{,\varphi_2}}{U},
1064: \label{varphiS}
1065: \end{eqnarray}
1066: %%%%%%%%%%%%%
1067: %%%%%%%%%%%%%
1068: \begin{eqnarray}
1069: -\frac{\dot{H}}{H^2}=\frac{1}{2\kappa^2}
1070: \left[\left(\frac{U_{,\varphi_1}}{U} \right)^2+
1071: e^{2F}\left(\frac{U_{,\varphi_2}}{U} \right)^2 \right].
1072: \label{dothubble}
1073: \end{eqnarray}
1074: %%%%%%%%%%%%%
1075: 
1076: In JBD and higher-dimensional theories where the potentials take the form,
1077: $U(\varphi_1, \varphi_2)= e^{-\beta \kappa \varphi_1}V(\varphi_2)$,
1078: it is straightforward to show that the scale factor evolves as power-law.
1079: %%jy-216a-start%%
1080: In this case integrating $\dot{\varphi}_1=\beta H/\kappa$ over $t$, we find
1081: %%%%%%%%%%%%%
1082: \begin{eqnarray}
1083: \varphi_1(t)=\frac{\beta}{\kappa} {\rm ln} \frac{a(t)}{a_f}+\varphi_{1f}
1084: \equiv -\frac{\beta}{\kappa}z+\varphi_{1f},
1085: \label{backphi1}
1086: \end{eqnarray}
1087: %%%%%%%%%%%%%
1088: where a subscript $f$ denotes the value of each quantity at the end of
1089: inflation and $z$ is the number of $e$-folds of inflationary expansion
1090: after the time $t$.
1091: %%jy-216a-end%%
1092: Assuming that $V(\varphi_2)$
1093: takes a constant value $V_0$ during inflation, one finds
1094: %%%%%%%%%%%%%
1095: \begin{eqnarray}
1096: a=a_0 \left[ \left\{ \frac{\kappa^2}{3}e^{-\beta\kappa\varphi_1(0)}
1097: V_0\right\}^{1/2} \frac{\beta^2}{2} (t-t_0) \right]^{2/\beta^2},
1098: \label{powerlaw}
1099: \end{eqnarray}
1100: %%%%%%%%%%%%%
1101: where $a_0$ and $t_0$ are constants.
1102: Then we have a power-law inflationary solution when $\beta<\sqrt{2}$.
1103: For example, in the JBD case with potential, (\ref{brans2}), inflation
1104: takes place with a large power exponent because $\beta$ is constrained
1105: to be $\beta=\sqrt{8/2\omega+3}~\lsim~0.034$ from observations
1106: \cite{omega}. In higher-dimensional theories with potential
1107: (\ref{FU2}), inflation is realized for arbitrary extra dimensions $D$,
1108: because the condition, $\beta<\sqrt{2}$, always holds.
1109: 
1110: Let us consider large scale perturbations with $k \ll aH$.  Neglecting
1111: $\dot{\Phi}$ and those terms which include second order time
1112: derivatives in Eqs.~(\ref{perturb1})-(\ref{perturb3}), one finds
1113: %%%%%%%%%%%%%
1114: \begin{eqnarray}
1115: \Phi=\frac{\kappa^2}{2H}
1116: \left( \dot{\varphi}_1 \delta \varphi_1
1117: +e^{-2F}\dot{\varphi}_2 \delta \varphi_2 \right),
1118: \label{perturb1slow}
1119: \end{eqnarray}
1120: %%%%%%%%%%%%%
1121: %%%%%%%%%%%%%
1122: \begin{eqnarray}
1123: 3H\delta\dot{\varphi}_1+U_{,\varphi_1\varphi_1}
1124: \delta\varphi_1
1125: +U_{,\varphi_1\varphi_2}\delta\varphi_2
1126: +2U_{,\varphi_1}\Phi=0,
1127: \label{perturb2slow}
1128: \end{eqnarray}
1129: %%%%%%%%%%%%%
1130: %%%%%%%%%%%%%
1131: \begin{eqnarray}
1132: 3H\delta\dot{\varphi}_2+\left(e^{2F}U_{,\varphi_2}\right)_{,\varphi_1}
1133: \delta\varphi_1+\left(e^{2F}U_{,\varphi_2}\right)_{,\varphi_2}
1134: \delta\varphi_2
1135: +2e^{2F}U_{,\varphi_2}\Phi=0.
1136: \label{perturb3slow}
1137: \end{eqnarray}
1138: %%%%%%%%%%%%%
1139: Note that this approximation may not be always valid especially when
1140: perturbations exhibit nonadiabatic growth during inflation.  We will check
1141: its validity in the next section.
1142: 
1143: In the slow-roll approximation, the expression (\ref{phiadia}) for the
1144: growing adiabatic mode is simplified:
1145: \begin{equation}
1146: \Phi=\Psi=h{\dot H\over 2H^2}~,~~\delta\varphi_n=-h{\dot\varphi_n\over 2H}~.
1147: \end{equation}
1148: To find a generic solution, we introduce new variables, $x$ and $y$, with
1149: $\delta\varphi_1=U_{,\varphi_1}x$ and $\delta \varphi_2=
1150: e^{2F}U_{,\varphi_2}y$. Then
1151: Eqs.~(\ref{perturb2slow}) and (\ref{perturb3slow}) yield
1152: %%%%%%%%%%%%%
1153: \begin{eqnarray}
1154: 3H\dot{x}+\frac{U_{,\varphi_1\varphi_2} U_{,\varphi_2}e^{2F}}{U_{,\varphi_1}}
1155: (y-x)+2\Phi=0,
1156: \label{x}
1157: \end{eqnarray}
1158: %%%%%%%%%%%%%
1159: %%%%%%%%%%%%%
1160: \begin{eqnarray}
1161: 3H\dot{y}+\frac{(e^{2F}U_{,\varphi_2})_{,\varphi_1} U_{,\varphi_1}e^{-2F}}
1162: {U_{,\varphi_2}}(x-y)+2\Phi=0.
1163: \label{y}
1164: \end{eqnarray}
1165: %%%%%%%%%%%%%
1166: Subtracting Eq.~(\ref{x}) from Eq.~(\ref{y}), we obtain the following
1167: integrated solution
1168: %%%%%%%%%%%%%
1169: \begin{eqnarray}
1170: y-x=Q_3\exp \left[\int \frac{A}{3H}dt\right],
1171: \label{y_x}
1172: \end{eqnarray}
1173: %%%%%%%%%%%%%
1174: where $Q_3$ is a constant, and
1175: %%%%%%%%%%%%%
1176: \begin{eqnarray}
1177: A \equiv \frac{U_{,\varphi_1\varphi_2} U_{,\varphi_2}e^{2F}}{U_{,\varphi_1}}
1178: +\frac{(e^{2F}U_{,\varphi_2})_{,\varphi_1}U_{,\varphi_1}e^{-2F}}
1179: {U_{,\varphi_2}}.
1180: \label{A}
1181: \end{eqnarray}
1182: %%%%%%%%%%%%%
1183: Taking notice of the relation
1184: %%%%%%%%%%%%%
1185: \begin{eqnarray}
1186: \Phi=\frac{\kappa^2}{2H}
1187: \left[ \dot{U}x+U_{,\varphi_2}\dot{\varphi}_2(y-x)\right]
1188: =\frac{\kappa^2}{2H}
1189: \left[ \dot{U}y+U_{,\varphi_1}\dot{\varphi}_1(x-y)\right],
1190: \label{Phislow}
1191: \end{eqnarray}
1192: %%%%%%%%%%%%%
1193: and making use of Eq.~(\ref{y_x}) and background equations
1194: (\ref{hubbleslow})-(\ref{varphi2slow}),
1195: we find
1196: %%%%%%%%%%%%%
1197: \begin{eqnarray}
1198: x &=& -\frac{Q_3}{U} \int \left[ \frac{H}{\kappa^2}
1199: \frac{U_{,\varphi_1\varphi_2} U_{,\varphi_2}e^{2F}}{U_{,\varphi_1}}
1200: +U_{,\varphi_2}\dot{\varphi}_2 \right] \frac{J}{U}dt, \\
1201: y &=& +\frac{Q_3}{U} \int \left[ \frac{H}{\kappa^2}
1202: \frac{(e^{2F}U_{,\varphi_2})_{,\varphi_1} U_{,\varphi_1}}{U_{,\varphi_2}}
1203: +U_{,\varphi_1}\dot{\varphi}_1 \right] \frac{J}{U}dt,
1204: \label{xandy}
1205: \end{eqnarray}
1206: %%%%%%%%%%%%%
1207: where
1208: %%%%%%%%%%%%%
1209: \begin{eqnarray}
1210: J \equiv U\exp \left[\int \frac{A}{3H}dt\right].
1211: \label{G}
1212: \end{eqnarray}
1213: %%%%%%%%%%%%%
1214: Then the final closed-form solutions for long-wave perturbations are
1215: expressed as
1216: %%%%%%%%%%%%%
1217: \begin{eqnarray}
1218: \delta \varphi_1 = ( {\rm ln} U)_{,\varphi_1}
1219: \left[Q_1+Q_3 \int_{t_*}^{t} \left({\rm ln} ({\rm ln} U)_{,\varphi_1}
1220: \right)_ {,\varphi_2} J d\varphi_2  \right],
1221: \label{finalslow1}
1222: \end{eqnarray}
1223: %%%%%%%%%%%%%
1224: %%%%%%%%%%%%%
1225: \begin{eqnarray}
1226: \delta \varphi_2 = e^{2F} ( {\rm ln} U)_{,\varphi_2}
1227: \left[Q_2-Q_3 \int_{t_*}^{t} \left({\rm ln} \left( e^{2F}({\rm ln} U)_
1228: {,\varphi_2} \right) \right)_{,\varphi_1}
1229:  J d\varphi_1  \right],
1230:  \label{finalslow2}
1231:  \end{eqnarray}
1232: %%%%%%%%%%%%%
1233: %%%%%%%%%%%%%
1234: \begin{eqnarray}
1235:  \Phi = -\frac12 \left[({\rm ln} U)_{,\varphi_1} \delta\varphi_1+
1236:  ({\rm ln} U)_{,\varphi_2} \delta\varphi_2 \right],
1237: \label{finalslow3}
1238: \end{eqnarray}
1239: %%%%%%%%%%%%%
1240: with
1241: %%%%%%%%%%%%%
1242: \begin{eqnarray}
1243: J=\exp \left\{ -\int_{t_*}^{t} \left[  \left({\rm ln}
1244: \left( e^{2F}({\rm ln} U)_{,\varphi_2} \right) \right)_{,\varphi_1} d\varphi_1+
1245: \left({\rm ln} ({\rm ln} U)_{,\varphi_1}
1246: \right)_ {,\varphi_2} d\varphi_2 \right] \right\}.
1247: \label{G2}
1248: \end{eqnarray}
1249: %%%%%%%%%%%%%
1250: Here integration constants, $Q_1$, $Q_2$, and $Q_3$ satisfy the relation
1251: $Q_2=Q_1+Q_3$, which comes from  Eq.~(\ref{y_x}).
1252: The $Q_3$ terms appear due to the presence of
1253: isocurvature perturbations.
1254: These constants are evaluated by the amplitudes
1255: of quantum fluctuations of scalar fields at horizon crossing, $t_*$.
1256: The fluctuations are generated by small scale perturbations ($k>aH$), so
1257: that they can be considered as free massless scalar fields which are
1258: described by independent random variables \cite{SY,PS}.
1259: Then the field
1260: perturbations when they crossed the Hubble radius ($k \simeq aH$)
1261: are written in the form:
1262: %%ST-2/17-start (change \epsilon to e)%%
1263: %%%%%%%%%%%%%
1264: \begin{eqnarray}
1265: \delta \varphi_1({\bf k})|_{t=t_*}=\frac{H(t_*)}{\sqrt{2k^3}}
1266: e_{\varphi_1}({\bf k}),~~~~
1267: \delta \varphi_2({\bf k})|_{t=t_*}=\frac{H(t_*)}{\sqrt{2k^3}}
1268: e^{F(t_*)} e_{\varphi_2}({\bf k}).
1269: \label{quantum}
1270: \end{eqnarray}
1271: %%%%%%%%%%%%%
1272: Here $e_{\varphi_1}$ and $e_{\varphi_2}$ are classical stochastic Gaussian
1273: quantities, described by
1274: %%%%%%%%%%%%%
1275: \begin{eqnarray}
1276: \langle e_{\varphi_1} ({\bf k}) \rangle =
1277: \langle e_{\varphi_2} ({\bf k}) \rangle=0,~~~~
1278: \langle e_{\varphi_1} ({\bf k}) e^*_{\varphi_2}({\bf k'})
1279: \rangle=\delta_{ij}\delta^{(3)}({\bf k}-{\bf k'}),
1280: \label{epsilon}
1281: \end{eqnarray}
1282: %%%%%%%%%%%%%
1283: where $i, j=\varphi_1, \varphi_2$.
1284: {}From Eqs.~(\ref{varphiS}), (\ref{finalslow1}), and (\ref{finalslow2}),
1285: we find
1286: %%%%%%%%%%%%%
1287: \begin{eqnarray}
1288: \frac{\delta \varphi_1}{\dot{\varphi}_1}=-\frac{\kappa^2}{H}
1289: \left[Q_1+Q_3 \int_{t_*}^{t} \left({\rm ln} ({\rm ln} U)_{,\varphi_1}
1290: \right)_ {,\varphi_2} J d\varphi_2  \right],
1291: \label{ratioslow1}
1292: \end{eqnarray}
1293: %%%%%%%%%%%%%
1294: %%%%%%%%%%%%%
1295: \begin{eqnarray}
1296: \frac{\delta \varphi_2}{\dot{\varphi}_2}=-\frac{\kappa^2}{H}
1297: \left[Q_1+Q_3-Q_3 \int_{t_*}^{t} \left({\rm ln} \left( e^{2F}({\rm ln} U)_
1298: {,\varphi_2} \right) \right)_{,\varphi_1}
1299:  J d\varphi_1  \right].
1300:  \label{ratioslow2}
1301:  \end{eqnarray}
1302: %%%%%%%%%%%%%
1303: Making use of Eqs.~(\ref{quantum}), (\ref{ratioslow1}), and
1304: (\ref{ratioslow2}), the integration constants are expressed as
1305: %%%%%%%%%%%%%
1306: \begin{eqnarray}
1307: Q_1=-\frac{H^2(t_*)}{\kappa^2\sqrt{2k^3}}
1308: \left( \frac{e_{\varphi_1}({\bf k})}
1309: {\dot{\varphi}_1}\right)_{t_*},~~~~
1310: Q_3=\frac{H^2(t_*)}{\kappa^2\sqrt{2k^3}} \left( \frac{e_{\varphi_1}({\bf k})}
1311: {\dot{\varphi}_1}- e^{F}\frac{e_{\varphi_2}({\bf k})}{\dot{\varphi}_2}
1312: \right)_{t_*}.
1313: \label{Qvalue}
1314: \end{eqnarray}
1315: In the next section, we will investigate the evolution of large-scale
1316: perturbations in specific gravity theories.
1317: 
1318: %%%%%%%%%%%%%%%%%%%%%%%%%%
1319: \section{Applications to specific gravity theories}
1320: %%%%%%%%%%%%%%%%%%%%%%%%%%
1321: 
1322: Among the generalized Einstein theories which we presented in Sec.~II,
1323: most of them take the following separated form of potentials except for the
1324: $f(R)$ theories:
1325: %%%%%%%%%%%%%
1326: \begin{eqnarray}
1327: U(\varphi_1, \varphi_2)=V_1(\varphi_1)V_2(\varphi_2).
1328: \label{poten}
1329: \end{eqnarray}
1330: %%%%%%%%%%%%%
1331: In this case, we have an integration constant by making use of
1332: Eq.~(\ref{varphiS}):
1333: %%jy-216b-start%%typos in the denominators corrected%%%
1334: \begin{eqnarray}
1335: C=-\kappa^2\int \frac{V_1}{V_{1,\varphi_1}}e^{2F}d\varphi_1+
1336: \kappa^2\int \frac{V_2}{V_{2,\varphi_2}} d\varphi_2.
1337: \label{C}
1338: \end{eqnarray}
1339: %%jy-216b-end%%%%%%%%%%%
1340: This characterizes the trajectory in field space in two-field inflation.
1341: 
1342: Since $J=e^{-2(F-F_*)}$ in the separated potential, (\ref{poten}),
1343: Eqs.~(\ref{finalslow1}) and (\ref{finalslow2}) are easily integrated to
1344: give
1345: %%%%%%%%%%%%%
1346: \begin{eqnarray}
1347: \delta\varphi_1=\frac{V_1'}{V_1} Q_1,~~~~~~
1348: \delta\varphi_2=\frac{V_2'}{V_2}
1349: \left[ Q_1 e^{2F}+Q_3 e^{2F_*} \right],
1350: \label{SEPfield}
1351: \end{eqnarray}
1352: %%%%%%%%%%%%%
1353: together with the gravitational potential,
1354: %%%%%%%%%%%%%
1355: \begin{eqnarray}
1356: \Phi=-\frac12 \left(\frac{V_1'}{V_1}\right)^2 
1357: Q_1-\frac12 \left(\frac{V_2'}{V_2}\right)^2
1358: \left( Q_1 e^{2F}+Q_3 e^{2F_*} \right).
1359: \label{SEPmet}
1360: \end{eqnarray}
1361: %%%%%%%%%%%%%
1362: %%ST-2/17-start (one more decomposition is added)%%
1363: Introducing new integration constants,
1364: $C_1 \equiv -\kappa^2(Q_1+Q_3e^{2F_*})$ and
1365: $C_3 \equiv -\kappa^2Q_3e^{2F_*}$, and making use of
1366: Eq.~(\ref{dothubble}), one finds
1367: %%ST-3/1-start
1368: %%%%%%%%%%%%%
1369: \begin{eqnarray}
1370: \delta \varphi_1=-(C_1-C_3)\frac{V_1'}{\kappa^2V_1},~~~~~~~
1371: \delta \varphi_2=-\left[C_1e^{2F}-C_3(e^{2F}-1) \right]
1372: \frac{V_2'}{\kappa^2V_2},
1373: \label{fieldper}
1374: \end{eqnarray}
1375: %%%%%%%%%%%%%
1376: %%%%%%%%%%%%%
1377: \begin{eqnarray}
1378: \Phi =C_1 \left( \epsilon_{\varphi_1}+e^{2F}\epsilon_{\varphi_2} 
1379: \right)-C_3 \left[\epsilon_{\varphi_1}+ 
1380: (e^{2F}-1) \epsilon_{\varphi_2} \right]
1381: =-C_1\frac{\dot{H}}{H^2}-C_3
1382: \left[\epsilon_{\varphi_1}+(e^{2F}-1) \epsilon_{\varphi_2} \right],
1383: \label{SEPmet2}
1384: \end{eqnarray}
1385: %%%%%%%%%%%%%
1386: %%ST-3/1-end
1387: where $\epsilon_{\varphi_1}$ and $\epsilon_{\varphi_2}$ are given by
1388: %%%%%%%%%%%%%
1389: \begin{eqnarray}
1390: \epsilon_{\varphi_1} \equiv \frac{1}{2\kappa^2}
1391: \left(\frac{V_1'}{V_1}\right)^2,~~~~
1392: \epsilon_{\varphi_2} \equiv \frac{1}{2\kappa^2}
1393: \left(\frac{V_2'}{V_2}\right)^2.
1394: \label{slowpara}
1395: \end{eqnarray}
1396: %%%%%%%%%%%%%
1397: {}From Eq.~(\ref{Qvalue}) $C_1$ and $C_3$ are expressed as
1398: %%%%%%%%%%%%%
1399: \begin{eqnarray}
1400: C_1=\frac{H^2(t_*)}{\sqrt{2k^3}}
1401: \left[ (1-e^{2F})\frac{e_{\varphi_1}({\bf k})}{\dot{\varphi}_1}
1402: +e^{3F} \frac{e_{\varphi_2}({\bf k})}{\dot{\varphi}_2} \right]_ {t_*},
1403: ~~~~~C_3=\frac{H^2(t_*)}{\sqrt{2k^3}}
1404: \left[e^{3F}\frac{e_{\varphi_2}({\bf k})}{\dot{\varphi}_2}-
1405: e^{2F}\frac{e_{\varphi_1}({\bf k})}{\dot{\varphi}_1} \right]_{t_*}.
1406: \label{NewC}
1407: \end{eqnarray}
1408: %%jy-3/12 start%%%%%%%%%%%
1409: The gravitational potential can also be decomposed in a different way.
1410: For example, let us introduce the integration constants $\tilde{C}_1$ and
1411: $\tilde{C}_3$, defined by
1412: %%jy-3/12 end%%%%%%%%%%%
1413: \begin{eqnarray}
1414: \tilde{C}_1 &\equiv& -\kappa^2 Q_1-\frac{\kappa^2 e^{2(F_*-F_f)}}
1415: {1+\alpha_f}Q_3=\frac{H^2(t_*)}{\sqrt{2k^3}}
1416: \left[\left(1-\frac{e^{2(F-F_f)}}{1+\alpha_f} \right)
1417: \frac{e_{\varphi_1}({\bf k})}{\dot{\varphi}_1}
1418: + \frac{e^{3F-2F_f}}{1+\alpha_f}
1419: \frac{e_{\varphi_2}({\bf k})} {\dot{\varphi}_2} \right]_ {t_*},
1420: \nonumber \\
1421: \tilde{C}_3 &\equiv& -\kappa^2 e^{2(F_*-F_f)}Q_3=
1422: \frac{H^2(t_*)}{\sqrt{2k^3}} \left[ e^{3F-2F_f}
1423: \frac{e_{\varphi_2}({\bf k})}{\dot{\varphi}_2}-
1424: e^{2(F-F_f)} \frac{e_{\varphi_1}({\bf k})}{\dot{\varphi}_1}
1425: \right]_ {t_*},
1426: \label{tildeC}
1427: \end{eqnarray}
1428: %%%%%%%%%%%%%
1429: where the subscript $f$ denotes the value at the end of inflation, and
1430: %%%%%%%%%%%%%
1431: \begin{eqnarray}
1432: \alpha \equiv \frac{\epsilon_{\varphi_1}}
1433: {e^{2F}\epsilon_{\varphi_2}}.
1434: \label{alpha}
1435: \end{eqnarray}
1436: %%%%%%%%%%%%%
1437: Then Eq.~(\ref{SEPmet}) reads
1438: %%%%%%%%%%%%%
1439: \begin{eqnarray}
1440: \Phi=-\tilde{C}_1\frac{\dot{H}}{H^2}-\tilde{C}_3
1441: \left[ \frac{\epsilon_{\varphi_1}}{1+\alpha_f}+
1442: \left(\frac{e^{2F}}{1+\alpha_f}-e^{2F_f}\right)
1443: \epsilon_{\varphi_2} \right].
1444: \label{SEPmet3}
1445: \end{eqnarray}
1446: %%%%%%%%%%%%%
1447: This decomposition corresponds to the case where the second
1448: term in the rhs of Eq.~(\ref{SEPmet3}) vanishes at the end
1449: of inflation.
1450: 
1451: During slow-roll we have that
1452: $|\dot{H}/H^2| \ll 1$ in Eq.~(\ref{dothubble}), which yields from
1453: Eqs.~(\ref{calR}), (\ref{SEPmet2}), and (\ref{SEPmet3}) as
1454: %%%%%%%%%%%%%
1455: \begin{eqnarray}
1456: {\cal R} \simeq -\frac{H^2}{\dot{H}}\Phi
1457: &=&
1458: C_1-C_3 \frac{\epsilon_{\varphi_1}+(e^{2F}-1)
1459: \epsilon_{\varphi_2}}
1460: {\epsilon_{\varphi_1}+e^{2F}\epsilon_{\varphi_2}}
1461: \label{SEPzeta_m1}\\
1462: &=& \tilde{C_1}-\tilde{C_3} \frac{\epsilon_{\varphi_1}
1463: +\left\{e^{2F}-(1+\alpha_f) e^{2F_f} \right\}
1464: \epsilon_{\varphi_2}}
1465: {(1+\alpha_f)(\epsilon_{\varphi_1} +
1466: e^{2F}\epsilon_{\varphi_2})}.
1467: \label{SEPzeta}
1468: \end{eqnarray}
1469: %%%%%%%%%%%%%
1470: Note that both decompositions coincide each other in the limit
1471: $\alpha_f \to 0$ and $F_f \to 0$.  When $\alpha_f$ and  $F_f$
1472: are nonvanishing, the second term in Eq.~(\ref{SEPmet2}) gives
1473: contributions to the gravitational potential, $\Phi$.  When this term is
1474: negligible relative to $-C_1\dot{H}/H^2$
1475: %%ST-3/1-start
1476: during the whole stage of inflation,
1477: %%ST-3/1-end
1478: the first and second terms in Eq.~(\ref{SEPmet2})
1479: can be identified as adiabatic and isocurvature modes, respectively.
1480: However, in some generalized Einstein theories which we discuss
1481: in the following
1482: subsections, the final $\Phi$ is dominated by the second term in
1483: Eq.~(\ref{SEPmet2}).  In those cases we can no longer regard the second
1484: term as the isocurvature mode at the end of inflation.
1485: 
1486: The relation (\ref{SEPzeta_m1}) or (\ref{SEPzeta}) indicates
1487: that ${\cal R}$ is not generally conserved during inflation
1488: when both fields are evolving due to
1489: the presence of isocurvature perturbations.  The generalized entropy
1490: perturbations are written as
1491: %%%%%%%%%%%%%
1492: \begin{eqnarray}
1493: S_{\varphi_1\varphi_2}=-C_3e^{-2F}=-\tilde{C}_3e^{2(F_f-F)}.
1494: \label{ent2}
1495: \end{eqnarray}
1496: %%%%%%%%%%%%%
1497: Since $S_{\varphi_1\varphi_2}$ and $Z$ do not vanish generally, these
1498: work as a source term for the change of ${\cal R}$. The evolution of
1499: ${\cal R}$ depends on specific gravity theories as we will show below.
1500: %%ST-2/17-end (one more decomposition is added)%%
1501: 
1502: 
1503: %
1504: \subsection{JBD theory}
1505: %
1506: 
1507: Let us first apply the results in the previous section to the JBD theory
1508: with Eq.~(\ref{brans2}).  In this case Eqs.~(\ref{SEPmet2}) and
1509: (\ref{SEPzeta_m1}) are
1510: %%%%%%%%%%%%%
1511: \begin{eqnarray}
1512: \Phi=-C_1\frac{\dot{H}}{H^2}-C_3\left[\frac{\beta^2}{2}+
1513: (e^{(\beta/2)\kappa\chi}-1) \epsilon_{\phi}\right],
1514: \label{JBDmetric}
1515: \end{eqnarray}
1516: %%%%%%%%%%%%%
1517: %%%%%%%%%%%%%
1518: \begin{eqnarray}
1519: {\cal R}=C_1-C_3\left[1-\frac{1}{e^{(\beta/2)\kappa\chi}
1520: +\beta^2/(2\epsilon_{\phi})}\right].
1521: \label{JBDzeta}
1522: \end{eqnarray}
1523: %%%%%%%%%%%%%
1524: 
1525: %%jy-216c-start%%%
1526: %%ST-2/18-start%%%
1527: In the JBD theory, $\beta$ is required to be $\beta~\lsim~0.034$ from
1528: observational constraints, which yields $\epsilon_{\chi}=\beta^2/2 \ll 1$.
1529: In addition to this, the value of $\chi$ should be practically vanishing,
1530: $\chi_f \simeq 0$, at the end of
1531: inflation in order to reproduce the present value of the gravitational
1532: constant, because $\chi$ generally grows only as logarithm of $t$ after
1533: inflation and is even constant during radiation-dominated stage in this
1534: theory.  These lead to $F_f=(\beta/4)\kappa \chi_f \simeq 0$
1535: and $\alpha_f=e^{-2F_f}\epsilon_{\chi_f}/\epsilon_{\phi_f} \simeq 0$,
1536: %%ST-3/1-start%%%
1537: which implies that the decomposition of Eq.~(\ref{SEPmet3}) looks almost
1538: the same form as that of (\ref{SEPmet2}).
1539: %%ST-3/1-end%%%
1540: Since the second term in the rhs
1541: of Eq.~(\ref{JBDmetric}) is negligible relative to the first term after
1542: inflation, the term proportional to $C_1$ represents the growing adiabatic
1543: mode \cite{MFB}, while the one proportional to $C_3$ corresponds to the
1544: isocurvature mode \cite{SY,CSY}.  During inflation, ${\cal R}$ evolves due to
1545: the change of the term,
1546: %%%%%%%%%%%%%
1547: \begin{eqnarray}
1548: W \equiv e^{(\beta/2)\kappa\chi}+\beta^2/(2\epsilon_{\phi}).
1549: \label{W}
1550: \end{eqnarray}
1551: %%%%%%%%%%%%%
1552: 
1553: As a specific model of inflation,
1554:  let us first consider chaotic inflation driven by a potential,
1555: $V(\phi)=\lambda_{2n}\phi^{2n}/(2n)$.
1556: In this model, inflation occurs at large $\phi$ and it is terminated
1557: when $|\dot{\phi}/\phi|$ becomes as large as $H$ at $\phi=\phi_f=
1558: \sqrt{2n}/\kappa$.
1559: Making use of Eqs.~(\ref{backphi1}) and (\ref{C}) we find
1560: %%%%%%%%%%%%%
1561: \begin{eqnarray}
1562: \kappa^2\int_{\phi_f}^{\phi}\frac{V(\varphi)}{V'(\varphi)}d\varphi
1563:  =\frac{\kappa^2}{4n}\lmk\phi^2-\frac{2n}{\kappa^2}\rmk
1564:  =  \frac{2}{\beta^2}(1-e^{\beta\kappa\chi/2})
1565:  = \frac{2}{\beta^2}(1-e^{-\beta^2z/2}) \simeq z.
1566: \end{eqnarray}
1567: The error in the last expression is less than $1.7\%$ for  $z \leq 60$ and
1568: $\beta \leq 0.034$.
1569: Then Eq.~(\ref{W}) is rewritten as
1570: %%%%%%%%%%%%%
1571: \begin{eqnarray}
1572: W=\left(1-\frac{2}{n}\right)e^{(\beta/2)\kappa\chi}
1573: +\frac{2}{n}+\frac{\beta^2}{2n}.
1574: \label{W2}
1575: \end{eqnarray}
1576: %%jy-216c-end%%%%%%%%%%%
1577: Since $W$ is constant for $n=2$, ${\cal R}$ is conserved
1578: in the quartic potential, $V(\phi)=\lambda_4 \phi^4/4$.
1579: In other cases such as the quadratic potential, ${\cal R}$ evolves
1580: during inflation due to the presence of isocurvature perturbations,
1581: although its change is typically small.
1582: At the end of inflation, ${\cal R}$ takes almost constant value,
1583: ${\cal R} \simeq C_1$.
1584: 
1585: %%%%%%%%%%
1586: \begin{figure}
1587: \begin{center}
1588: \singlefig{12cm}{Fig1.eps}
1589: \begin{figcaption}{Fig1}{12cm}
1590: The evolution of the adiabatic metric perturbation
1591: $\tilde{\Phi}_{\rm ad} \equiv k^{3/2} \Phi_{\rm ad}$,
1592: the isocurvature metric perturbation
1593: $\tilde{\Phi}_{\rm iso} \equiv k^{3/2} \Phi_{\rm iso}$,
1594: and the curvature perturbation $\tilde{\cal R} \equiv k^{3/2} 
1595: {\cal R}$ for the case of $\beta=0.09$
1596: in the JBD theory with quadratic inflaton potential,
1597: $V(\phi)=m^2\phi^2/2$.
1598: We choose the initial values of scalar fields as $\phi=3m_{\rm pl}$
1599: and $\chi=-1.2m_{\rm pl}$, in which case $\chi$ is nearly zero
1600: at the end of inflation.
1601: \end{figcaption}
1602: \end{center}
1603: \end{figure}
1604: %%%%%%%%%%
1605: 
1606: One may worry that the above results are obtained
1607: by imposing slow-roll conditions, which are only approximations to the
1608: full equations of motion.
1609: In order to answer such suspicions, we numerically solved full equations
1610: (\ref{perturb4})-(\ref{perturb3}) along with the background equations
1611: (\ref{hubble1})-(\ref{varphi2}).
1612: We adopt the quadratic potential $V(\phi)=m^2\phi^2/2$,
1613: and start integrating from
1614: about  60 e-folds before the end of inflation.
1615: We found that the evolution of field and metric perturbations are well
1616: described by analytic estimations except around the end of inflation.  The
1617: evolution of $\Phi$ for the adiabatic and isocurvature mode is plotted in
1618: Fig.~\ref{Fig1}, which shows that the
1619: contribution of the isocurvature mode is small as estimated by
1620: Eq.~(\ref{JBDmetric}).  The adiabatic growth of the gravitational potential
1621: terminates for $mt~\gsim~20$, after which the system enters a reheating
1622: stage.  During reheating no additional growth of super-Hubble metric
1623: perturbations occurs in this scenario, unless some interactions between
1624: inflaton and other field, $\sigma$, such as $g^2\phi^2\sigma^2/2$ are
1625: not introduced.
1626: 
1627: The evolution of ${\cal R}$ for $\beta=0.09$
1628: (corresponding to $\omega=500$) is plotted in Fig.~1,
1629: which shows that the change of ${\cal R}$ is small during inflation.
1630: We have also confirmed that the variation 
1631: of ${\cal R}$ is negligible in the case of $\beta~\lsim~0.034$.
1632: Since its growth is sourced by the $e^{(\beta/2)\kappa\chi}$ term in
1633: Eq.~(\ref{W2}), we find in Fig.~1 that ${\cal R}$ approaches 
1634: almost constant value around the end of inflation.  
1635: During reheating ${\cal R}$ is conserved
1636: except for the short period when $\dot{H}$ passes through zero.
1637: 
1638: %%jy-217a-start%%
1639: Thus one can use analytic expressions for adiabatic curvature
1640: perturbations based on slow-roll approximations in order to constrain
1641: the model parameters of the potential.
1642: Note that the number of e-folds, $\zk$, after the comoving wave-number $k$
1643: leaves
1644: the Hubble radius during inflation satisfies
1645: \beq
1646:   \frac{k}{k_f}=e^{-\lmk 1-\beta^2/2 \rmk\zk}
1647:   (2\zk)^{n/2},~~~~~\mbox{for}~~ 1 \ll \zk \lesssim 60.
1648: \eeq
1649: We can then express the amplitude of curvature perturbation on scale
1650: $l=2\pi/k$ as\cite{SY}
1651: \beqa \Phi(l)&=&\lkk 1+\frac{2}{3(1+w)}\rkk^{-1}
1652: \frac{\sqrt{2k^3\langle |C_1|^2 \rangle}}{2\pi} \nonumber \\
1653: &=&\lkk 1+\frac{2}{3(1+w)}\rkk^{-1}
1654: \frac{\kappa^2 e^{\beta^2\zk/2}}{2\pi}
1655: \sqrt{\frac{\lambda_{2n}}{6n}
1656: \lmk \frac{4n\zk}{\kappa^2}\rmk^{n}}
1657: \lkk e^{-\beta^2\zk/4}
1658: \sqrt{\frac{\zk}{n}}
1659: +\frac{1}{\beta}(1-e^{-\beta^2\zk/2}) \rkk,
1660: \eeqa
1661: which is valid until the second horizon crossing after inflation and
1662: also for $1 \ll \zk~\lsim~60$. Here $w$ is the ratio of the pressure to the
1663: energy density.
1664: 
1665: Since the large-angular-scale anisotropy of background radiation due to
1666: the Sachs-Wolfe effect is given by $\delta T/T = \Phi/3$,
1667: we can normalize the value of $\lambda_{2n}$ by the COBE
1668: normalization \cite{LR99}.  For $\beta=0.034$, since this gives the relation,
1669: \beqa 
1670: \frac{\delta T}{T} &=& \frac{1}{3}\Phi (\zk \simeq 60) \simeq \lnk
1671: \begin{array}{lll} 9 \lambda_2/m_{\rm pl}
1672:  \equiv & 9 m/m_{\rm pl} &
1673: \mbox{~~~~~for $n=1$}\\
1674:   28 \sqrt{\lambda_4}&~ &
1675:   \mbox{~~~~~for $n=2$}\\
1676:   \end{array} \right. \\
1677:   &=& 1.1\times 10^{-5},
1678: \eeqa
1679: one finds
1680: \beqa m&=& 1\times 10^{13}~\mbox{GeV},~~~~~n=1, \\
1681:   \lambda_4 &=& 2\times 10^{-13},~~~~~~~~~~n=2,
1682: \eeqa
1683: which is not much different from the values obtained assuming the Einstein
1684: gravity \cite{Salopek}.
1685: Since the behavior of the system approaches to that in the
1686: Einstein gravity as we increase $\omega$, we can conclude that in
1687: Brans-Dicke theory, model
1688: parameters of chaotic inflation should take the same value as
1689: in the Einstein gravity.
1690: 
1691: Next we consider new inflation with a potential
1692: \beq
1693:   V(\phi)=V_0 -\frac{\lambda}{4}\phi^4,
1694: \eeq
1695: for which we find,
1696: \beq
1697:   \kappa^2\int_{\phi_f}^{\phi}\frac{V(\varphi)}{V'(\varphi)}d\varphi
1698: \cong \frac{\kappa^2V_0}{2\lambda}\lmk \frac{1}{\phi^2}
1699: - \frac{1}{\phi_f^2}\rmk \cong \frac{\kappa^2}{2\lambda\phi^2}
1700: \simeq z.
1701: \eeq
1702: We can again express the amplitude of curvature fluctuation as a
1703: function of $\zk$, which is now related with $k$
1704: as,
1705: \beq
1706: \frac{k}{k_f}=e^{-\lmk 1-\beta^2/2 \rmk\zk},
1707: \eeq
1708: \beq
1709: \Phi(l)=\lkk 1+\frac{2}{3(1+w)}\rkk^{-1}
1710: \lkk e^{\beta^2\zk/4}
1711: \sqrt{\frac{\lambda}{3}}(2\zk)^{3/2} +
1712: \kappa H_f \frac{e^{\beta^2\zk/2}}{\beta}
1713: (1-e^{-\beta^2\zk/2})\rkk,
1714: \eeq
1715: with $H_f\equiv \sqrt{\kappa^2V_0/3}$.
1716: Taking $\beta=0.034$ again, it predicts the amplitude of
1717: $\delta T/T$ to be compared with COBE data as
1718: \beq
1719: \frac{\delta T}{T}(\zk \simeq 60)
1720: \simeq 30 \sqrt{\lambda} +0.21 \frac{H_f}{\mpl}.  \label{newinfdT}
1721: \eeq
1722: Since $H_f$ should  also satisfy
1723: \beq
1724: \frac{H_f}{m_{\rm pl}}~\lsim~10^{-5},
1725: \eeq
1726: to suppress long-wave gravitational
1727: radiation of quantum origin \cite{GW}, we obtain
1728: \beq
1729: \lambda~\lsim~1\times 10^{-13},
1730: \eeq
1731: {}from Eq.~(\ref{newinfdT}).  Again its amplitude is
1732: practically no different from the case of the Einstein gravity.
1733: %%jy-217a-end%%%
1734: %%ST-2/18-end%%%
1735: 
1736: %
1737: \subsection{Non-minimally coupled scalar field case}
1738: %
1739: 
1740: %%ST-3/1-start%%%
1741: Let us first briefly review the single-field inflationary scenario
1742: with a non-minimally coupled scalar field ($\xi R\phi^2/2)$.
1743: In chaotic inflationary models, Futamase and Maeda \cite{FM} found that the
1744: non-minimal
1745: coupling is constrained as $|\xi|~\lsim~10^{-3}$ in the quadratic
1746: potential, by the requirement of sufficient amount of
1747: inflation\footnote{This constraint is loosened by considering topological
1748: inflation, see \cite{SY99}.}.
1749: In the quartic potential, such a constraint is absent
1750: for negative $\xi$, and as a bonus, the fine tuning problem
1751: of the self-coupling $\lambda$ in the minimally coupled case
1752: can be relaxed by considering large negative values of
1753: $\xi$ \cite{Spo,SBB,FU}.
1754: Several authors evaluated scalar and tensor perturbations generated during
1755: inflation \cite{MS,Kaiser,Hwang99,KF} and preheating \cite{SB00} in this
1756: model.  Since the system is reduced to the single-field case with a
1757: modified inflaton potential by a conformal transformation, we cannot
1758: expect nonadiabatic growth of ${\cal R}$ on large scales.
1759: %%ST-3/1-end%%%
1760: 
1761: We shall proceed to the case of the non-minimally coupled $\psi$ field
1762: with Eq.~(\ref{nonmin3}) in the presence of inflaton, $\phi$.
1763: In this theory the evolution of field and metric perturbations was studied in
1764: \cite{TY} in the Jordan frame.  It was found that ${\cal R}$ can grow
1765: nonadiabatically during inflation on super-horizon scales for negative $\xi$.
1766: Here we will show that similar results are obtained by the analysis in the
1767: Einstein frame.
1768: 
1769: {}From Eqs.~(\ref{fieldper}), (\ref{SEPmet2}),
1770: and (\ref{SEPzeta_m1}), we
1771: obtain the following explicit solutions:
1772: %%%%%%%%%%%%%
1773: \begin{eqnarray}
1774: \delta\chi=-4(C_1-C_3) \frac{\xi\psi}
1775: {\sqrt{1-(1-6\xi)\xi\kappa^2\psi^2}},~~~~
1776: \delta\phi=-\frac{V'(\phi)}{\kappa^2V(\phi)} \left[C_1
1777: (1-\xi\kappa^2\psi^2)+C_3 (\xi\kappa^2\psi^2) \right],
1778: \label{NONfield2}
1779: \end{eqnarray}
1780: %%%%%%%%%%%%%
1781: %%%%%%%%%%%%%
1782: \begin{eqnarray}
1783: \Phi = -C_1\frac{\dot{H}}{H^2}-C_3\left[
1784: \epsilon_{\psi}- (\xi\kappa^2\psi^2) \epsilon_{\phi} \right],
1785: \label{NONmet2}
1786: \end{eqnarray}
1787: %%%%%%%%%%%%%
1788: %%%%%%%%%%%%%
1789: \begin{eqnarray}
1790: {\cal R}= C_1-C_3 \frac{\epsilon_{\psi}- (\xi\kappa^2\psi^2)
1791: \epsilon_{\phi}} {\epsilon_{\psi}+(1-\xi\kappa^2\psi^2
1792: )\epsilon_{\phi}},
1793: \label{NONzeta2}
1794: \end{eqnarray}
1795: %%%%%%%%%%%%%
1796: with
1797: %%%%%%%%%%%%%
1798: \begin{eqnarray}
1799: \epsilon_{\psi}=\frac{8(\xi\kappa\psi)^2} {1-(1-6\xi)\xi\kappa^2\psi^2},~~~~
1800: \epsilon_{\phi}=\frac{1}{2\kappa^2} \left(\frac{V'(\phi)}{V(\phi)}\right)^2.
1801: \label{NONepsilon}
1802: \end{eqnarray}
1803: %%%%%%%%%%%%%
1804: When $\xi$ is negative, the coefficient of $\psi$ in the rhs
1805: of Eq.~(\ref{varphiS}) is always positive,
1806: which leads to the rapid growth of $\psi$
1807: (and $\chi$).  Eq.~(\ref{NONfield2}) indicates that long wave $\delta\chi$
1808: fluctuations are amplified with $|\psi|$ being increased.  This is due to the
1809: fact that the effective mass of $\delta\chi$ becomes negative after horizon
1810: crossing \cite{TY}, whose property is different from the JBD case.  In the
1811: JBD case, $\delta\chi$ is almost constant during
1812: inflation [see Eq.~(\ref{fieldper}) with $V_1=e^{-\beta\kappa\chi}$],
1813: which restricts the nonadiabatic growth of large scale metric perturbations.
1814: In contrast, in the present model, $\Phi$ and ${\cal R}$ exhibit strong
1815: amplification due to the excitation of low momentum field perturbations
1816: unless $|\psi|$ is initially very small.
1817: 
1818: The second terms in Eqs.~(\ref{NONmet2}) and (\ref{NONzeta2}) appear
1819: in the presence of non-minimal coupling, whose contributions are
1820: negligible when $|\xi|\kappa^2\psi^2 \ll 1$.
1821: With the increase of $|\psi|$, however,
1822: isocurvature perturbations are generated during inflation, which can lead
1823: to nonadiabatic growth of $\Phi$ and ${\cal R}$.  When the second term in
1824: Eq.~(\ref{NONmet2}) grows to of order the first term, the adiabatic mode
1825: includes the isocurvature mode partially.  In this case one cannot
1826: completely decompose adiabatic and isocurvature modes in the final results
1827: by the expression, Eq.~(\ref{NONmet2}).
1828: 
1829: 
1830: %%ST-2/18-start (change initial \chi to coincide with figs)%%%
1831: Let us consider the massive chaotic inflationary scenario
1832: with initial conditions, $\phi_*=3m_{\rm pl}$
1833: and $\chi_*=10^{-2}m_{\rm pl}$.
1834: We plot in Fig.~2 the evolution of field perturbations for $\xi=-0.02$
1835: in two cases, i.e., (i) solving directly the perturbed equations,
1836: (\ref{perturb4})-(\ref{perturb3}), (ii) using the slow-roll solutions,
1837: Eq.~(\ref{NONfield2}).  In spite of the rapid growth of field perturbations,
1838: slow-roll analysis agrees reasonably well with full numerical results.
1839: The enhancement of $\delta\chi_k$ fluctuations stimulates the amplification
1840: of $\delta\phi_k$ fluctuations for $mt~\gsim~10$.  Inflationary period
1841: ends around $mt \simeq 17$, after which the slow-roll results begin to
1842: fail.  In Fig.~3, the evolution of $\Phi$ and ${\cal R}$ is depicted.  
1843: We also plot the first and the second term in the rhs of 
1844: Eq.~(\ref{NONmet2}), where
1845: we denote $\Phi_1^{(s)}$ and $\Phi_2^{(s)}$, respectively. 
1846: Although $\Phi$ is dominated by the $\Phi_1^{(s)}$ term in the initial stage,
1847: $\Phi_2^{(s)}$ catches up $\Phi_1^{(s)}$ around $mt \simeq 7$ , 
1848: after which the $\Phi_2^{(s)}$ term completely determines 
1849: the evolution of $\Phi$.  We
1850: find in Fig.~3 that slow-roll approximations are valid right up until the
1851: end of inflation.  The growth of metric perturbations stops when the system
1852: enters a reheating stage.
1853: 
1854: %%ST-3/1-start
1855: If we use the decomposition of Eq.~(\ref{SEPmet3}), the second term gives
1856: negligible contribution to the gravitational potential around the end of
1857: inflation.  In the early stage of inflation, however, its contribution is
1858: comparable to the first term, in which case the isocurvature mode cannot
1859: be completely separated from the adiabatic mode.  When fluctuations are
1860: sufficiently amplified, it is inevitable that both adiabatic and
1861: isocurvature modes mix each other with the growth of fluctuations, which
1862: means that complete decomposition is difficult during the whole
1863: inflationary stage.
1864: %%%%%%%%%%
1865: \begin{figure}
1866: \begin{center}
1867: \singlefig{12cm}{Fig2.eps}
1868: \begin{figcaption}{Fig2}{12cm}
1869: The evolution of background fields, $\tilde{\phi}=\phi/m_{\rm pl}$
1870: and $\tilde{\chi}=\chi/m_{\rm pl}$, and large-scale field perturbations,
1871: $\delta\tilde{\phi} \equiv k^{3/2} \delta\phi/m_{\rm pl}$
1872: and $\delta\tilde{\chi} \equiv k^{3/2} \delta\chi/m_{\rm pl}$,
1873: in the massive chaotic inflationary scenario
1874: with a non-minimally coupled $\chi$ field for $\xi=-0.02$.
1875: The initial values of scalar fields are chosen as
1876: $\phi_*=3m_{\rm pl}$ and $\chi_*=10^{-2}m_{\rm pl}$.
1877: The slow-roll results of Eq.~(\ref{NONfield2}) are also plotted,
1878: where we denote $\delta\tilde{\phi}^{(s)}$ and $\delta\tilde{\chi}^{(s)}$ in the
1879:  figure.
1880: We find that slow-roll approximations are valid except in the final stage of
1881: inflation ($mt>17$). The evolution of
1882: $\delta\tilde{\phi}^{(s)}$ and $\delta\tilde{\chi}^{(s)}$ 
1883: after inflation is not shown.
1884: \end{figcaption}
1885: \end{center}
1886: \end{figure}
1887: %%%%%%%%%%
1888: 
1889: %%%%%%%%%%
1890: \begin{figure}
1891: \begin{center}
1892: \singlefig{12cm}{Fig3.eps}
1893: \begin{figcaption}{Fig3}{12cm}
1894: The evolution of the metric perturbation
1895: $\tilde{\Phi} \equiv k^{3/2} \Phi$ and the curvature
1896: perturbation $\tilde{\cal R}\equiv k^{3/2} {\cal R}$
1897: with same parameters as in Fig.~2.
1898: The amplification of field perturbations leads to the nonadiabatic 
1899: growth of $\Phi$ and ${\cal R}$.
1900: We also plot $\tilde{\Phi}_1^{(s)} \equiv k^{3/2} \Phi_1^{(s)}$
1901: and $\tilde{\Phi}_2^{(s)} \equiv k^{3/2} \Phi_2^{(s)}$,
1902: where $\Phi_1^{(s)}$ and $\Phi_2^{(s)}$ denote the first and the second
1903: terms in Eq.~(\ref{NONmet2}), respectively.
1904: $\Phi$ is mainly sourced by the
1905: $\Phi_2^{(s)}$ term, after $\Phi_2^{(s)}$ 
1906: catches up $\Phi_1^{(s)}$.  The evolution of 
1907: $\Phi_1^{(s)}$ and $\Phi_2^{(s)}$ after inflation is 
1908: not shown.
1909: \end{figcaption}
1910: \end{center}
1911: \end{figure}
1912: %%%%%%%%%%
1913: 
1914: In Fig.~3 the curvature perturbation, ${\cal R}$, 
1915: is nonadiabatically amplified sourced by the second term 
1916: in Eq.~(\ref{NONzeta2}).  Whether this occurs or
1917: not depends upon the strength of the coupling, $\xi$, and the initial
1918: $\chi$.  When both are small and the second term in Eq.~(\ref{NONzeta2}) is
1919: negligible relative to the $C_1$ term during inflation, we can regard the
1920: first and second terms in Eq.~(\ref{NONzeta2}) as adiabatic and
1921: isocurvature modes, respectively.  In the simulations of Figs.~2 and 3, we
1922: take $\chi_*=10^{-2}m_{\rm pl}$, in which case numerical calculations
1923: imply that the conservation of ${\cal R}$ is violated for $\xi~\lsim~-0.01$.
1924: When $\xi~\lsim~-1$, strong amplification of ${\cal R}$ is inevitable 
1925: even for very small values of $\chi$ far less than $m_{\rm pl}$.
1926: For positive $\xi$, conservation of ${\cal R}$ is typically preserved due
1927: to an exponential suppression of $\chi$ during inflation \cite{TY,BV,BPTV}.
1928: Regarding detailed investigation about the observational constraints of the
1929: strength of $\xi$, see \cite{TY} whose results are similar to
1930: those in the analysis of the Einstein frame.
1931: %%ST-3/1-start
1932: %%ST-2/18-end (change initial \chi to coincide with figs)%%%
1933: 
1934: 
1935: %
1936: \subsection{Higher-dimensional theories}
1937: %
1938: 
1939: %%ST-2/19-start %%%
1940: In the higher-dimensional theory with Eq.~(\ref{FU2}), the kinetic term
1941: takes a canonical form.  In this theory the condition, $D>1$, gives the
1942: constraint, $\sqrt{2/3}<\beta<\sqrt{2}$, which is
1943: different form the JBD theory with $\beta \ll 1$.  Larger values of $\beta$
1944: correspond to the steep exponential potential of the $\chi$ field, which
1945: leads to the rapid evolution toward the $\chi$ direction.
1946: In this case inflaton decreases slowly relative to the $\chi$ field.
1947: Then the expansion of the
1948: universe is described by the power-law solution,
1949: Eq.~(\ref{powerlaw}).
1950: 
1951: For example, in the polynomial inflaton potential, $V(\phi)
1952: =\lambda_{2n}\phi^{2n}/(2n)$, classical trajectories of scalar fields
1953: are given by Eq.~(\ref{C}) as
1954: %%%%%%%%%%%%%
1955: \begin{eqnarray}
1956: C=\frac{\kappa}{\beta}\chi+\frac{\kappa^2}{4n}\phi^2.
1957: \label{trajectry}
1958: \end{eqnarray}
1959: %%%%%%%%%%%%%
1960: Differentiating Eq.~(\ref{trajectry}) with respect to $t$ yields
1961: %%%%%%%%%%%%%
1962: \begin{eqnarray}
1963: \left| \frac{\dot{\chi}}{\dot{\phi}} \right|=
1964: \frac{\beta\kappa}{2n}|\phi|=\frac{\sqrt{2\pi}\beta}{n}
1965: \left| \frac{\phi}{m_{\rm pl}} \right|.
1966: \label{dif}
1967: \end{eqnarray}
1968: %%%%%%%%%%%%%
1969: This relation indicates that for the values of $\phi$ greater than $m_{\rm pl}$
1970: with  $\beta$ and $n$ being of order unity, $|\dot{\chi}|$ is typically larger
1971: than
1972:  $|\dot{\phi}|$, in which case $\chi$ rapidly evolves along the exponential
1973: potential.
1974: The power-law inflation continues until $|\dot{\phi}|$ grows comparable
1975: to $|\dot{\chi}|$, corresponding to $\phi/m_{\rm pl} \simeq n/(\sqrt{2\pi}\beta)
1976: $.
1977: After $\phi$ falls down this value, $\phi$ begins to evolve faster than
1978: $\chi$ toward the local minimum at $\phi=0$ in the $\phi$ direction.
1979: In this stage the system deviates from the power-law expansion, (\ref{powerlaw}).
1980: 
1981: {}From Eqs.~(\ref{SEPmet2}), (\ref{SEPmet3}),
1982: (\ref{SEPzeta_m1}), and (\ref{SEPzeta}),
1983: $\Phi$ and ${\cal R}$ evolve during inflation as
1984: %%%%%%%%%%%%%
1985: \begin{eqnarray}
1986: \Phi =-C_1\frac{\dot{H}}{H^2}-C_3\frac{\beta^2}{2}
1987: =-\tilde{C}_1\frac{\dot{H}}{H^2}-\tilde{C}_3
1988: \frac{\beta^2-2\alpha_f\epsilon_{\phi}}{2(1+\alpha_f)},
1989: \label{highermet}
1990: \end{eqnarray}
1991: %%%%%%%%%%%%%
1992: %%%%%%%%%%%%%
1993: \begin{eqnarray}
1994: {\cal R} = C_1-C_3 \frac{\beta^2}{\beta^2+2\epsilon_{\phi}}
1995: =\tilde{C}_1-\tilde{C}_3 \frac{\beta^2-2\alpha_f\epsilon_{\phi}}
1996: {(1+\alpha_f)(\beta^2+2\epsilon_{\phi})},
1997: \label{higherzeta}
1998: \end{eqnarray}
1999: %%%%%%%%%%%%%
2000: where $\alpha_f=\beta^2/(2\epsilon_{\phi_f})$ is of order unity
2001: for $\sqrt{2/3}<\beta<\sqrt{2}$.
2002: Since the condition, $\beta^2 \gg \epsilon_{\phi}$,
2003: holds during power law-inflation, the
2004: curvature perturbation in this stage takes almost a constant value,
2005: ${\cal R} \simeq C_1-C_3=\tilde{C}_1-\tilde{C}_3/(1+\alpha_f)$.
2006: As $|\dot{\phi}|$ grows relative to $|\dot{\chi}|$,
2007: ${\cal R}$ begins to evolve due to the change of $\epsilon_{\phi}$
2008: in Eq.~(\ref{higherzeta}).
2009: This corresponds to the stage where
2010: deviations from power-law inflation become relevant.
2011: When $\epsilon_{\phi}$ in Eq.~(\ref{higherzeta}) becomes
2012: comparable to $\beta^2/2$ (i.e., $\alpha_f \simeq 1)$ around the
2013: end of inflation, we have ${\cal R} \simeq C_1-C_3 \alpha_f/(1+\alpha_f)
2014: =\tilde{C}_1$. After inflation ${\cal R}$ takes this conserved value.
2015: %%ST-2/19-end %%%
2016: 
2017: %
2018: \subsection{The $R^2$ theory with a non-minimally coupled
2019: $\chi$ field}
2020: %
2021: 
2022: In the $f(R)$ theories, effective potentials  do
2023: not generally take separated forms, Eq.~(\ref{poten}), as found
2024: in Eq.~(\ref{FU}).  Nevertheless we have closed form solutions,
2025: Eqs.~(\ref{finalslow1})-(\ref{G2}), by which the evolution of
2026: cosmological perturbations can be studied analytically.
2027: 
2028: Let us analyze the $R^2$ inflationary scenario with a
2029: non-minimally coupled $\chi$ field as one example of
2030: the $f(R)$ theory [see Eqs.~(\ref{R2FR}) and
2031: (\ref{R2potential})].
2032: Note that when $\chi=0$ the system has an effective potential,
2033: %%%%%%%%%%%%%
2034: \begin{eqnarray}
2035: U=\frac{m_{\rm pl}^4}{(32\pi)^2\ab}
2036: \left(1-e^{-(\sqrt{6}/3)\kappa\phi}\right)^2.
2037: \label{R2pot}
2038: \end{eqnarray}
2039: %%%%%%%%%%%%%
2040: The $\phi$ field defined by Eq.~(\ref{newphi}) plays the role of an inflaton
2041: and leads to inflationary expansion of the Universe in the region,
2042: $\phi~\gsim~m_{\rm pl}$ \cite{R2,TMT}. In the absence of the
2043: non-minimally coupled $\chi$ field, the resulting spectrum of density
2044: perturbations after the end of inflation was found in \cite{St83}
2045: (using equations for perturbations of the FRW model for the Einstein gravity
2046: with one-loop quantum corrections derived in \cite{St81}) and then
2047: rederived in \cite{density_R2}. Here we study how the effect of
2048: non-minimal coupling alters the adiabatic evolution of cosmological
2049: perturbations in the single field case.
2050: 
2051: In the presence of
2052: non-minimal coupling, Eq.~(\ref{G2}) is reduced to
2053: %%%%%%%%%%%%%
2054: \begin{eqnarray}
2055: J=\frac{[1-(1-\xi\kappa^2\chi^2)e^{-(\sqrt{6}/3)\kappa\phi}]
2056: (1-\xi\kappa^2\chi_*^2)}
2057: {[1-(1-\xi\kappa^2\chi_*^2)e^{-(\sqrt{6}/3)\kappa\phi_*}]
2058: (1-\xi\kappa^2\chi^2)}.
2059: \label{G3}
2060: \end{eqnarray}
2061: %%%%%%%%%%%%%
2062: Then Eqs.~(\ref{finalslow1}) and (\ref{finalslow2}) are integrated to give
2063: %%%%%%%%%%%%%
2064: \begin{eqnarray}
2065: \delta\phi=-\frac{2\sqrt{6}(1-\xi\kappa^2\chi^2)
2066: e^{-(\sqrt{6}/3)\kappa\phi}}{3\kappa
2067: [1-(1-\xi\kappa^2\chi^2) e^{-(\sqrt{6}/3)\kappa\phi}]}
2068: \left[C_1-C_3
2069: \frac{\xi\kappa^2(\chi^2-\chi_*^2)}
2070: {[1-(1-\xi\kappa^2\chi_*^2)e^{-(\sqrt{6}/3)\kappa\phi_*}]
2071: (1-\xi\kappa^2\chi^2)}\right],
2072: \label{R2field1}
2073: \end{eqnarray}
2074: %%%%%%%%%%%%%
2075: \begin{eqnarray}
2076: \delta\chi=-\frac{4\xi\chi}
2077: {1-(1-\xi\kappa^2\chi^2) e^{-(\sqrt{6}/3)\kappa\phi}}
2078: \left[C_1+C_3
2079: \frac{1-(1-\xi\kappa^2\chi_*^2)e^{-(\sqrt{6}/3)\kappa\phi}}
2080: {1-(1-\xi\kappa^2\chi_*^2)e^{-(\sqrt{6}/3)\kappa\phi_*}} \right],
2081: \label{R2field2}
2082: \end{eqnarray}
2083: %%%%%%%%%%%%%
2084: where $C_1=-\kappa^2Q_1$ and $C_3=-\kappa^2Q_3$.
2085: Therefore $\Phi$ and ${\cal R}$ are expressed as
2086: %%%%%%%%%%%%%
2087: \begin{eqnarray}
2088: \Phi = -C_1 \frac{\dot{H}}{H^2}-
2089: \frac{C_3}{1-(1-\xi\kappa^2\chi_*^2) e^{-(\sqrt{6}/3)\kappa\phi_*}}
2090: \left[ \epsilon_{\phi} \frac{\xi\kappa^2(\chi^2-\chi_*^2)}
2091: {1-\xi\kappa^2\chi^2}-\epsilon_{\chi}
2092: \left\{e^{(\sqrt{6}/3)\kappa\phi}-(1-\xi\kappa^2\chi_*^2)\right\}
2093: \right],
2094: \label{R2PHI}
2095: \end{eqnarray}
2096: %%%%%%%%%%%%%
2097: %%%%%%%%%%%%%
2098: \begin{eqnarray}
2099: {\cal R}= C_1-\frac{C_3}{1-(1-\xi\kappa^2\chi_*^2)
2100: e^{-(\sqrt{6}/3)\kappa\phi_*}}
2101: \frac{\epsilon_{\phi}~\xi\kappa^2(\chi^2-\chi_*^2)-
2102: \epsilon_{\chi}\left\{e^{(\sqrt{6}/3)\kappa\phi}-
2103: (1-\xi\kappa^2\chi_*^2)\right\}}
2104: {\epsilon_{\phi}+e^{(\sqrt{6}/3)\kappa\phi}\epsilon_{\chi}},
2105: \label{R2zeta}
2106: \end{eqnarray}
2107: %%%%%%%%%%%%%
2108: where $\epsilon_{\phi}$ and $\epsilon_{\chi}$ are defined by
2109: %%ST-3/2-start
2110: %%%%ST-2/18-start (add definitions)%%%%%%%%
2111: \begin{eqnarray}
2112: \epsilon_{\phi} &\equiv& \frac{1}{2\kappa^2}\left(
2113: \frac{U_{,\phi}}{U} \right)^2=
2114: \frac43 \left( \frac{(1-\xi\kappa^2\chi^2)
2115: e^{-(\sqrt{6}/3)\kappa\phi}}
2116: {1-(1-\xi\kappa^2\chi^2)e^{-(\sqrt{6}/3)\kappa\phi}}\right)^2,
2117: \\
2118: \epsilon_{\chi} &\equiv& \frac{1}{2\kappa^2}\left(
2119: \frac{U_{,\chi}}{U} \right)^2= 8 \left( \frac{\xi\kappa\chi
2120: e^{-(\sqrt{6}/3)\kappa\phi}}
2121: {1-(1-\xi\kappa^2\chi^2)e^{-(\sqrt{6}/3)\kappa\phi}}\right)^2.
2122: \label{R2slow}
2123: \end{eqnarray}
2124: %%%ST-2/18-start (add definitions)%%%%%%
2125: In the absence of non-minimal coupling, one finds adiabatic results,
2126: $\Phi=-C_1\dot{H}/{H^2}$ and ${\cal R}=C_1$.
2127: In two-field inflation with a non-minimally coupled
2128: $\chi$ field, the presence of isocurvature perturbations can lead to
2129: nonadiabatic growth of $\Phi$ and ${\cal R}$ as found
2130: in the second terms in Eqs.~(\ref{R2PHI}) and (\ref{R2zeta}).
2131: Their contributions are negligible when the conditions,
2132: $|\xi|\kappa^2\chi^2 \ll 1$ and $\epsilon_{\chi} \ll 1$,
2133: holds during inflation.
2134: The latter condition is similar to the former one
2135: when $|\xi|~\lsim~1$.
2136: 
2137: {}From Eq.~(\ref{varphiS}), since  $\dot{\chi}$ is
2138: approximately written as
2139: %%%%%%%%%%%%%
2140: \begin{eqnarray}
2141: \dot{\chi}=-\frac{4\xi H}{1-(1-\xi\kappa^2\chi^2)
2142: e^{-(\sqrt{6}/3)\kappa\phi}} \chi,
2143: \label{R2dotchi}
2144: \end{eqnarray}
2145: %%%%%%%%%%%%%
2146: the $\chi$ field exhibits exponential decrease for positive values of $\xi$.
2147: For negative $\xi$, however, $\chi$ is exponentially amplified
2148: during inflation, which means that the condition,
2149: $|\xi|\kappa^2\chi^2 \ll 1$, can be violated.
2150: The long-wave $\delta\chi$ fluctuation grows with the increase of $\chi$
2151: as found in Eq.~(\ref{R2field2}).
2152: On the other hand, the growth of $\delta\phi$ begins only when
2153: the second term in Eq.~(\ref{R2field1})
2154: becomes comparable to the first term.
2155: 
2156: %%%%%%%%%%
2157: \begin{figure}
2158: \begin{center}
2159: \singlefig{12cm}{Fig4.eps}
2160: \begin{figcaption}{Fig4}{12cm}
2161: The evolution of background fields, $\tilde{\phi}=\phi/m_{\rm pl}$
2162: and $\tilde{\chi}=\chi/m_{\rm pl}$, and long-wave field perturbations,
2163: $\delta\tilde{\phi} \equiv k^{3/2} \delta\phi/m_{\rm pl}$ and
2164: $\delta\tilde{\chi} \equiv k^{3/2} \delta\chi/m_{\rm pl}$ as
2165: a function of time, $\bar{t} \equiv m_{\rm pl}t/\sqrt{96\pi \ab}$
2166: in the $R^2$ inflationary scenario with a non-minimally coupled
2167: $\chi$ field for $\xi=-0.025$.  The initial values of scalar fields are chosen
2168: as
2169: $\phi_*=1.1m_{\rm pl}$ and $\chi_*=10^{-3}m_{\rm pl}$.  We also plot the
2170: slow-roll results, where we denote $\delta\tilde{\phi}^{(s)}$ and
2171: $\delta\tilde{\chi}^{(s)}$ in the figure.  The evolution of
2172: $\delta\tilde{\phi}^{(s)}$ and $\delta\tilde{\chi}^{(s)}$ 
2173: after inflation is not shown.
2174: \end{figcaption}
2175: \end{center}
2176: \end{figure}
2177: %%%%%%%%%%
2178: 
2179: 
2180: %%%%%%%%%%
2181: \begin{figure}
2182: \begin{center}
2183: \singlefig{12cm}{Fig5.eps}
2184: \begin{figcaption}{Fig5}{12cm}
2185: The evolution of the metric perturbation $\tilde{\Phi} \equiv k^{3/2}\Phi$
2186: and the curvature perturbation $\tilde{\cal R} \equiv k^{3/2} {\cal R}$
2187: in the $R^2$ inflationary scenario with
2188: same parameters as in Fig.~4.
2189: We also plot $\tilde{\Phi}_1^{(s)} \equiv k^{3/2} \Phi_1^{(s)}$
2190: and $\tilde{\Phi}_2^{(s)} \equiv k^{3/2} \Phi_2^{(s)}$,
2191: where $\Phi_1^{(s)}$ and $\Phi_2^{(s)}$ denote the first and the second
2192: terms in Eq.~(\ref{R2PHI}), respectively.
2193: The evolution of $\Phi_1^{(s)}$ and $\Phi_2^{(s)}$ 
2194: after inflation is not shown.
2195: \end{figcaption}
2196: \end{center}
2197: \end{figure}
2198: %%%%%%%%%%
2199: 
2200: We plot the evolution of $\delta\chi$, $\delta\phi$, and the slow-roll
2201: results (\ref{R2field1})  and (\ref{R2field2})  for $\xi=-0.025$
2202: with initial conditions,
2203: $\phi_*=1.1m_{\rm pl}$ and $\chi_*=10^{-3}m_{\rm pl}$.
2204: In this case inflation ends around $\tilde{t}\equiv
2205: m_{\rm pl}t/\sqrt{96\pi \ab} \approx 130$ with e-foldings, $N \approx 63$.
2206: Again the slow-roll analysis is quite reliable except around
2207: the end of inflation.
2208: 
2209: %%ST-2/18-start (change initial \chi to coincide with figs)%%%
2210: In Fig.~5 we also depict the evolution of $\Phi$ and ${\cal R}$,
2211: and the first ($=\Phi_1^{(s)}$) and second
2212: ($=\Phi_2^{(s)}$) terms in Eq.~(\ref{R2PHI}).  In the initial stage of
2213: inflation where $\chi$ and $\delta\chi$ are not sufficiently amplified, the
2214: gravitational potential is dominated by the $\Phi_1^{(s)}$ term, in which
2215: case $\Phi_2^{(s)}$ may be regarded as the isocurvature mode.  However,
2216: after $\tilde{t} \approx 70$ where $\Phi_2^{(s)}$ catches up
2217: $\Phi_1^{(s)}$, we find in Fig.~5 that $\Phi_2^{(s)}$
2218: mainly contributes to the
2219: gravitational potential.  As is similar to the case of the chaotic
2220: inflationary scenario with a non-minimally coupled $\chi$ field,
2221: adiabatic and isocurvature modes mix each other
2222: with the growth of the $\chi$ fluctuation.
2223: If one defines the isocurvature mode as the one which gives
2224: negligible contribution to the gravitational potential, it can not be
2225: completely separated from the adiabatic mode during the whole stage of
2226: inflation.
2227: 
2228: The conservation of ${\cal R}$ is typically violated when the  term
2229: proportional to $C_3$ in Eq.~(\ref{R2zeta}) surpasses the one
2230: proportional to $C_1$.  For the initial values,
2231: $\phi_*=1.1m_{\rm pl}$ and $\chi_*=10^{-3}m_{\rm pl}$,
2232: ${\cal R}$ exhibits nonadiabatic growth for $\xi~\lsim~-0.02$.
2233: Negative large non-minimal coupling such as $\xi~\lsim~-1$ leads to strong
2234: amplification of ${\cal R}$ unless $\chi$ is initially very small.
2235: Although we do not make detailed analysis here,
2236: the basic property is quite similar to the case of the subsection B.
2237: These results are also expected to hold for other inflationary models with a
2238: non-minimally coupled $\chi$ field, since the scalar curvature is
2239: proportional to the potential energy of inflaton which slowly decreases
2240: during inflation.
2241: %%ST-2/18-end(change initial \chi to coincide with figs)%%%
2242: %%ST-3/2-end
2243: 
2244: 
2245: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2246:  \section{Conclusions}
2247: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2248: 
2249: %%%%%aas%%%%%%%%%%%%%
2250: We have studied generation and evolution of adiabatic and
2251: isocurvature perturbations during multi-field inflation in
2252: generalized Einstein theories. Most of the generalized Einstein
2253: theories recast to the Lagrangian (\ref{lagrangian}) in the
2254: Einstein frame by conformal transformations. While behavior of
2255: adiabatic perturbations is universal and is given by Eqs. (\ref{adia},
2256: \ref{phiadia}) in the long-wave limit ($k\to 0$), isocurvature
2257: perturbations behave differently depending on specific gravity
2258: theories. Making use of slow-roll approximations, we have obtained
2259: closed form solutions for all non-decaying field and metric perturbations
2260: in the long-wave limit. The existence of isocurvature perturbations
2261: may lead to significant variations of the curvature perturbations
2262: ${\cal R}$ and $\zeta$.
2263: %%%%%%%%aas%%%%%%%%%%%%%%%
2264: 
2265: In this work we considered the following four gravity theories.
2266: 
2267: (1) The Jordan-Brans-Dicke theory with a Brans-Dicke field $\chi$ and
2268: inflaton $\phi$.  Using the Brans-Dicke parameter constrained by
2269: observations, the isocurvature mode in the gravitational potential $\Phi$ is
2270: negligible relative to the adiabatic mode.  Therefore the variation of
2271: ${\cal R}$ is typically small in this theory.
2272: %%jy-3/12 start%%%
2273: In particular, for the quartic potential, ${\cal R}$ is conserved in the
2274: slow-roll analysis.
2275: %%jy-3/12 end%%%%%
2276: 
2277: (2) A non-minimally coupled scalar field $\chi$ in the presence of inflaton
2278: $\phi$.  When the coupling $\xi$ is negative, $\chi$ and its long-wave
2279: fluctuations exhibit exponential increase during inflation,
2280: leading to the nonadiabatic amplification of $\Phi$ and ${\cal R}$
2281: due to the existence of isocurvature perturbations.
2282: Even in this case we find
2283: that slow-roll analysis agrees well with full numerical results.
2284: When field and metric perturbations are sufficiently amplified,
2285: adiabatic and isocurvature modes of the gravitational potential mix
2286: each other.
2287: 
2288: (3) Higher-dimensional Kaluza-Klein theory with dilaton $\chi$ and
2289: inflaton $\phi$.  In this theory the inflationary period can be divided
2290: %%jy-3/12 start %%%%%
2291: into  two stages: the first is the power-law inflationary stage
2292: where $\chi$ evolves along the exponential potential and
2293: the second is the deviation from power-law inflation
2294: due to the rapid evolution of $\phi$ around the end
2295: of inflation.
2296: %%jy-3/12 end%%%%%
2297: In the former stage ${\cal R}$ is nearly constant but its change
2298: occurs at the transition between two stages.
2299: 
2300: (4) $R^2$ theory with a non-minimally coupled scalar field $\chi$.
2301: This system has an additional scalar field $\phi$ playing the role of
2302: inflaton after conformal transformations.  Although this theory has
2303: a coupled effective potential which is different from the above theories
2304: (1)-(3), we have integrated forms of long-wave field and metric
2305: perturbations by slow-roll analysis, which is found to be quite reliable
2306: right up until the end of inflation.
2307: Negative non-minimal coupling again leads
2308: to the nonadiabatic growth of $\Phi$ and ${\cal R}$, in which case
2309: complete decomposition between adiabatic and isocurvature modes
2310: is difficult.
2311: 
2312: While we analyzed generalized Einstein theories involving
2313: two scalar fields,
2314: there exist other multi-field inflationary scenarios such as hybrid
2315: inflation \cite{hybrid} and models of two interacting
2316: scalar fields \cite{KL87}. During preheating after inflation,
2317: there has been growing interest
2318: about the evolution of cosmological perturbations for the simple
2319: two-field model with potential
2320: $U(\phi,\chi)=\lambda \phi^4/4+
2321: g^2\phi^2\chi^2/2$ \cite{BV,selfpre}.
2322: It is certainly of interest to constrain realistic multi-field
2323: inflationary models based on particle physics using density perturbations
2324: generated during inflation, together with constraints by gravitinos
2325: \cite{gravitinos} and possible primordial black hole
2326: over-production during preheating \cite{PBH}.
2327: %%ST-3/3-end
2328: 
2329: 
2330: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2331: \section*{ACKNOWLEDGEMENTS}
2332: AS is thankful to Profs. Katsuhiko Sato and Masahiro Kawasaki for
2333: hospitality in RESCEU, the University of Tokyo. He was also partially
2334: supported by the Russian Fund for Fundamental Research, grants 99-02-16224
2335: and 00-15-96699.
2336: ST thanks Bruce A. Bassett, Kei-ichi Maeda, Naoshi Sugiyama,
2337: Atsushi Taruya, Takashi Torii, Hiroki Yajima, 
2338: and David Wands for useful discussions.
2339: He is also thankful for financial support from the JSPS (No. 04942).
2340: The work of JY was partially supported by the Monbukagakusho
2341: Grant-in-Aid, Priority Area ``Supersymmetry and Unified Theory of
2342: Elementary Particles''(\#707) and the Monbukagakusho Grant-in-Aid for
2343: Scientific Research No.\ 112440063.
2344: 
2345: 
2346: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2347: \begin{thebibliography}{99}
2348: 
2349: \bibitem{inf}
2350: E.W. Kolb and M.S. Turner, {\it The Early Universe}
2351: (Addison-Wesley, Redwood City, California, 1990);
2352: A.D. Linde, {\it Particle Physics and Inflationary Cosmology}
2353: (Harwood, Chur, Switzerland, 1990);
2354: A.R. Liddle and D.H. Lyth, {\it Cosmological inflation and
2355: Large-Scale Structure} (Cambridge University Press, 2000).
2356: 
2357: \bibitem{oriinf}
2358: A.H. Guth, Phys. Rev. D {\bf 23}, 347 (1981);
2359: K. Sato, Mon. Not. R. Astr. Soc. {\bf 195}, 467 (1981).
2360: 
2361: \bibitem{newinf}
2362: A.D. Linde, Phys. Lett. {\bf B108}, 389 (1982);
2363: A. Albrecht and P.J. Steinhardt,
2364: Phys. Rev. Lett. {\bf 48}, 1220 (1982).
2365: 
2366: \bibitem{chaoinf}
2367: A.D. Linde, Phys. Lett. {\bf B129}, 177 (1983).
2368: 
2369: \bibitem{R2}
2370: A.A. Starobinsky, Phys. Lett. {\bf B91}, 99 (1980).
2371: 
2372: \bibitem{W84}
2373: B. Whitt, Phys. Lett. {\bf B145}, 176 (1984).
2374: 
2375: \bibitem{Got92}
2376: S. Gottl\"ober, V. M\"uller, H.-J. Schmidt, and A.A. Starobinsky,
2377: Int. J. Mod. Phys. {\bf D1}, 257 (1992).
2378: 
2379: \bibitem{St79}
2380: A.A. Starobinsky, JETP Lett. {\bf 30}, 682 (1979).
2381: 
2382: \bibitem{pert}
2383: S.W. Hawking, Phys. Lett. {\bf B115}, 295 (1982);
2384: A.A. Starobinsky, Phys. Lett. {\bf B117}, 175 (1982);
2385: A.H. Guth and S.Y. Pi,
2386: Phys. Rev. Lett. {\bf 49}, 1110 (1982).
2387: 
2388: \bibitem{L80}
2389: V.N. Lukash, Sov. Phys. JETP {\bf 52}, 807 (1980).
2390: 
2391: \bibitem{M81}
2392: V.F. Mukhanov and G.V. Chibisov, JETP Lett. {\bf  33}, 532 (1981).
2393: 
2394: \bibitem{MSYY}
2395: H. Murayama, H. Suzuki, T. Yanagida, and J. Yokoyama,
2396: Phys. Rev. Lett. {\bf 70}, 1912 (1993);
2397: Phys. Rev. D {\bf50}, R2356 (1994).
2398: 
2399: \bibitem{LR99}
2400: D.H. Lyth and A. Riotto, Phys. Rept. {\bf 314}, 1 (1999).
2401: 
2402: \bibitem{extinf}
2403: D. La and P.J. Steinhardt, Phys. Rev. Lett. {\bf 62}, 376 (1989);
2404: P.J. Steinhardt and  F.S. Accetta,
2405: Phys. Rev. Lett. {\bf 64}, 2470 (1990).
2406: 
2407: \bibitem{BD}
2408: C. Brans and R.H. Dicke, Phys. Rev. {\bf 24}, 925 (1961).
2409: 
2410: \bibitem{presc}
2411: D. La, P.J. Steinhardt, and E.W. Bertschinger, Phys. Lett. {\bf B231},
2412: 231 (1989); E. Weinberg, Phys. Rev. D {\bf 40}, 3950 (1989).
2413: 
2414: \bibitem{soft}
2415: A.L. Berkin, K. Maeda, and J. Yokoyama, Phys. Rev. Lett. {\bf 65},
2416: 141 (1990).
2417: 
2418: \bibitem{Lin90}
2419: A.D. Linde, Phys. Lett. {\bf B238}, 160 (1990).
2420: 
2421: \bibitem{FT85}
2422: E.S. Fradkin and A.A. Tseytlin, Phys. Lett. {\bf B158}, 316 (1985);
2423: Nucl. Phys. {\bf B261}, 1 (1985).
2424: 
2425: \bibitem{C85}
2426: C.G. Callan, D. Friedan, E.J. Martinec, and M.J. Perry,
2427: Nucl. Phys. {\bf B262}, 593 (1985);
2428: C.G. Callan, I.R. Klebanov, and M.J. Perry, Nucl. Phys. {\bf B278},
2429: 78 (1986).
2430: 
2431: \bibitem{BM}
2432: A.L. Berkin and K. Maeda, Phys. Rev. D {\bf 44}, 1691 (1991).
2433: 
2434: \bibitem{Mc}
2435: J. McDonald, Phys. Rev. D {\bf 44}, 2314 (1991).
2436: 
2437: \bibitem{MM}
2438: S. Mollerach and S. Matarrese, Phys. Rev. D {\bf 45}, 1961 (1992).
2439: 
2440: \bibitem{Der}
2441: N. Deruelle, C. Gundlach, and D. Langlois,
2442: Phys. Rev. {\bf D 46}, 5337 (1992).
2443: 
2444: \bibitem{Garcia}
2445: J. Garc\'{\i}a-Bellido, A.D. Linde and D.A. Linde,
2446: Phys. Rev. D {\bf 50}, 730 (1994);
2447: J. Garc\'{\i}a-Bellido, Nucl. Phys. {\bf B423}, 221 (1994).
2448: 
2449: \bibitem{BST}
2450: J.M. Bardeen, P.J. Steinhardt, and M.S. Turner, Phys. Rev. D {\bf 28},
2451: 679 (1983).
2452: 
2453: \bibitem{Lyth85}
2454: D.H. Lyth, Phys. Rev. D {\bf 31}, 1792 (1985).
2455: 
2456: \bibitem{KS}
2457: H. Kodama and M. Sasaki, Prog. Theor. Phys. Suppl. {\bf 78}, 1 (1984).
2458: 
2459: \bibitem{SY}
2460: A.A. Starobinsky and J. Yokoyama, in:
2461: Proc. of The Fourth Workshop on General Relativity and Gravitation,
2462: ed. by K. Nakao {\it et al.} (Kyoto University, 1994), p. 381
2463: (gr-qc/9502002).
2464: 
2465: \bibitem{St85}
2466: A.A. Starobinsky, JETP Lett. {\bf 42}, 152 (1985).
2467: 
2468: \bibitem{PS}
2469: D. Polarski and A.A. Starobinsky, Nucl. Phys. {\bf B385}, 623 (1992);
2470: Phys. Rev. D {\bf 50}, 6123 (1994).
2471: 
2472: \bibitem{DN}
2473: T. Damour and K. Nordtvedt, Phys. Rev. Lett. {\bf 70}, 2217 (1993);
2474: Phys. Rev. D {\bf 48}, 3436 (1993).
2475: 
2476: \bibitem{GW1}
2477: J. Garc\'{\i}a-Bellido and D. Wands, Phys. Rev. D {\bf 52}, 6739
2478: (1995).
2479: 
2480: \bibitem{CSY}
2481: T. Chiba, N. Sugiyama, and J. Yokoyama,
2482: Nucl. Phys. B {\bf 530}, 304 (1998).
2483: 
2484: \bibitem{SS}
2485: M. Sasaki and E. Stewart, Prog. Theor. Phys. {\bf 95}, 71 (1996).
2486: 
2487: \bibitem{GW2}
2488: J. Garc\'{\i}a-Bellido and D. Wands, Phys. Rev. D {\bf 53}, 437
2489: (1996).
2490: 
2491: \bibitem{multiMS}
2492: V.F. Mukhanov and P.J. Steinhardt, Phys. Lett. {\bf B422}, 52 (1998).
2493: 
2494: \bibitem{LH60}
2495: E.M. Lifshitz and I.M. Khalatnikov, Zh. Eksp. Teor. Fiz. {\bf 39},
2496: 149 (1960).
2497: 
2498: \bibitem{LH63}
2499: E.M. Lifshitz and I.M. Khalatnikov, Adv. of Phys. {\bf 12}, 208
2500: (1963).
2501: 
2502: \bibitem{KH98}
2503: H. Kodama and T. Hamazaki, Phys. Rev. D {\bf 57}, 7177 (1998).
2504: 
2505: \bibitem{ST98}
2506: M. Sasaki and T. Tanaka, Prog. Theor. Phys. {\bf 99}, 763 (1998).
2507: 
2508: \bibitem{NT98}
2509: Y. Nambu and A. Taruya, Class. Quant. Grav. {\bf 15}, 2761 (1998).
2510: 
2511: \bibitem{WMLL}
2512: D. Wands, K.A. Malik, D.H. Lyth and A.R. Liddle, Phys. Rev. D
2513: {\bf 62}, 043527 (2000).
2514: 
2515: \bibitem{Hwang}
2516: J. Hwang, Class. Quantum Grav. {\bf 7}, 1613 (1990); Phys. Rev. D
2517: {\bf 42}, 2601 (1990); Astroph. J. {\bf 375}, 443 (1991); Phys. Rev.
2518: D {\bf 53}, 762 (1996).
2519: 
2520: \bibitem{Maeda89}
2521: K. Maeda, Phys. Rev. D {\bf 39}, 3159 (1989).
2522: 
2523: \bibitem{omega}
2524: C.M. Will, Living. Rev. Rel. {\bf 4}, 1 (2001) (gr-qc/0103036).
2525: 
2526: \bibitem{TY}
2527: S. Tsujikawa and H. Yajima, Phys. Rev. D {\bf 62},
2528: 123512 (2000).
2529: 
2530: \bibitem{SH}
2531: V. Sahni and S. Habib, Phys. Rev. Lett. {\bf 81}, 1766 (1998);
2532: see also B.A. Bassett ans S. Liberati, Phys. Rev. D {\bf 58},
2533: 021302 (1998); S. Tsujikawa, K. Maeda, and T. Torii,
2534: Phys. Rev. D {\bf 60}, 063515 (1999).
2535: 
2536: \bibitem{induced}
2537: A. Zee, Phys. Rev. Lett. {\bf 42}, 417 (1979);
2538: S.L. Adler, Phys. Lett. {\bf B95}, 241 (1980);
2539: F.S. Accetta, D. J. Zoller, and M.S. Turner,  Phys. Rev. D {\bf
2540: 31},  3046 (1985).
2541: 
2542: \bibitem{CW}
2543: P. Candelas and S. Weinberg, Nucl. Phys. B {\bf 237}, 397 (1984).
2544: 
2545: \bibitem{Amendola}
2546: L. Amendola,  E.W. Kolb, M. Litterio, and F. Occhionero,
2547: Phys. Rev. D {\bf 42}, 1944 (1990).
2548: 
2549: \bibitem{shinji00}
2550: S. Tsujikawa, JHEP {\bf 0007}, 024 (2000).
2551: 
2552: \bibitem{Boom01}
2553: C.B. Netterfield {\it et al.}, astro-ph/0104460 (2001).
2554: 
2555: \bibitem{Bardeen}
2556: J.M. Bardeen, Phys. Rev. D {\bf 22}, 1882 (1980).
2557: 
2558: \bibitem{MFB}
2559: V.F. Mukhanov, H.A. Feldman, and R.H. Brandenberger,
2560: Phys. Rep. {\bf 215}, 293 (1992).
2561: 
2562: \bibitem{LL93}
2563: A.R. Liddle and D.H. Lyth, Phys. Rep. {\bf 231}, 1 (1993).
2564: 
2565: \bibitem{MSvar}
2566: V.F. Mukhanov, JETP Lett. {\bf 41}, 493 (1985);
2567: M. Sasaki, Prog. Theor. Phys. {\bf 76}, 1036 (1986).
2568: 
2569: \bibitem{Leach}
2570: S.M. Leach, M. Sasaki, D. Wands, and A.R. Liddle, 
2571: Phys. Rev. D {\bf 64}, 023512 (2001).
2572: 
2573: \bibitem{GWBM}
2574: C. Gordon, D. Wands, B.A. Bassett, and R. Maartens,
2575: Phys. Rev. D {\bf 63}, 023506 (2001).
2576: 
2577: \bibitem{HN}
2578: J. Hwang and H. Noh, Phys. Lett. {\bf B495}, 277 (2000).
2579: 
2580: \bibitem{Salopek}
2581: D.S. Salopek, Phys. Rev. Lett. {\bf 69}, 3602 (1992).
2582: 
2583: \bibitem{GW}
2584: V.A. Rubakov, M.V. Sazhin, and A.V. Veryaskin, Phys. Lett. {\bf B115},
2585: 189 (1982).
2586: 
2587: \bibitem{FM}
2588: T. Futamase and K. Maeda, Phys. Rev. D {\bf 39}, 399 (1989).
2589: 
2590: \bibitem{SY99}
2591: N. Sakai and J. Yokoyama, Phys. Lett. {\bf B456}, 113 (1999).
2592: 
2593: \bibitem{Spo}
2594: B.L. Spokoiny, Phys. Lett. {\bf B147}, 39 (1984).
2595: 
2596: \bibitem{SBB}
2597: D.S. Salopek, J.R. Bond, and J.M. Bardeen,
2598: Phys. Rev. D {\bf 40}, 1753 (1989).
2599: 
2600: \bibitem{FU}
2601: R. Fakir and W.G. Unruh, Phys. Rev. D {\bf 41}, 1783 (1990).
2602: 
2603: \bibitem{MS}
2604: N. Makino and M. Sasaki, Prog. Theor. Phys. {\bf 86}, 103 (1991).
2605: 
2606: \bibitem{Kaiser}
2607: D.I. Kaiser, Phys. Rev. D {\bf 52}, 4295 (1995).
2608: 
2609: \bibitem{Hwang99}
2610: J. Hwang and H. Noh, Phys. Rev. D {\bf 60}, 123001 (1999).
2611: 
2612: \bibitem{KF}
2613: E. Komatsu and T. Futamase, Phys. Rev. D {\bf 58},
2614: 023004 (1998); Phys. Rev. D {\bf 59}, 064029 (1999).
2615: 
2616: \bibitem{SB00}
2617: S. Tsujikawa and B.A. Bassett, Phys. Rev. D {\bf 62}, 043510 (2000);
2618: S. Tsujikawa, K. Maeda, and T. Torii, Phys. Rev. D {\bf 61}, 103501
2619: (2000).
2620: 
2621: \bibitem{BV}
2622: B.A. Bassett and F. Viniegra, Phys. Rev. D {\bf 62}, 043507 (2000).
2623: 
2624: \bibitem{BPTV}
2625: B.A. Bassett, G. Pollifrone, S. Tsujikawa, and F. Viniegra,
2626: Phys. Rev. D {\bf 63}, 103515  (2001).
2627: 
2628: \bibitem{TMT}
2629: S. Tsujikawa, K. Maeda, and T. Torii,
2630: Phys. Rev. D {\bf 60}, 123505 (1999).
2631: 
2632: \bibitem{St83}
2633: A.A. Starobinsky, Sov. Astron. Lett. {\bf 9}, 302 (1983).
2634: 
2635: \bibitem{St81}
2636: A.A. Starobinsky, JETP Lett. {\bf 34}, 438 (1981).
2637: 
2638: \bibitem{density_R2}
2639: L.A. Kofman and V.F. Mukhanov, JETP Lett. {\bf 44},
2640: 619 (1986);  L.A. Kofman, V.F. Mukhanov,
2641: and D.Yu. Pogosyan, Sov. Phys. JETP {\bf 66}, 433 (1987);
2642: L.A. Kofman, V.F. Mukhanov, and D.Yu. Pogosyan,
2643: Phys. Lett. {\bf B193}, 427 (1987).
2644: 
2645: \bibitem{hybrid}
2646: A.D. Linde, Phys. Lett. {\bf B259}, 38 (1991); Phys. Rev.  D {\bf 49},
2647: 748 (1994); E.J. Copeland, A.R. Liddle, D.H. Lyth, E.D. Stewart, and
2648: D. Wands, Phys.  Rev.  D {\bf 49}, 6410 (1994).
2649: 
2650: \bibitem{KL87}
2651: L.A. Kofman and A.D. Linde, Nucl. Phys. {\bf B282}, 555 (1987);
2652: L.A. Kofman and D.Yu. Pogosyan, Phys. Lett. {\bf B214}, 508 (1988).
2653: 
2654: \bibitem{selfpre}
2655: F. Finelli and R.H. Brandenberger, Phys. Rev. D {\bf 62},
2656: 083502 (2000); S. Tsujikawa, B.A. Bassett, and F. Viniegra,
2657: JHEP {\bf 0008}, 019 (2000);
2658: J.P.  Zibin, R.H. Brandenberger, and D. Scott, Phys. Rev. D {\bf 63},
2659: 043511 (2001); F. Finelli and S. Khlebnikov, 
2660: Phys. Lett. {\bf B504}, 309 (2001); hep-ph/0107143.
2661: 
2662: \bibitem{gravitinos}
2663: A.L. Maroto and  A. Mazumdar, Phys. Rev. Lett. {\bf 84}, 1655 (2000);
2664: R. Kallosh, L. Kofman, A. Linde, and A. Van Proeyen,
2665: Phys. Rev. D {\bf 61}, 103503 (2000);
2666: G.F. Giudice, A. Riotto and  I.I. Tkachev,
2667: JHEP {\bf 9911}, 036 (1999);
2668: D.H. Lyth and H.B. Kim, hep-ph/0011262;
2669: H.P. Nilles, M. Peloso, and L. Sorbo,
2670: hep-ph/0102264; JHEP {\bf 0104}, 004 (2001).
2671: 
2672: \bibitem{PBH}
2673: A.M. Green and K.A. Malik, Phys. Rev. D {\bf 64}, 021301 (2001); 
2674: B.A. Bassett and
2675: S. Tsujikawa, Phys. Rev. D {\bf 63}, 123503 (2001).
2676: 
2677: 
2678: \end{thebibliography}
2679: \end{document}
2680: 
2681: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2682: 
2683: %%%%%%%%%%%%%%%%%%%%
2684: %%%  Figures  %%%%%
2685: %%%%%%%%%%%%%%%%%%%%
2686: 
2687: