hep-th0308057/v2.tex
1: \documentclass[12pt]{article}
2: \input epsf
3: \usepackage{epsfig%,showkeys
4: }
5: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6: \def\ie{{\it i.e.}}
7: \def\IZ{\relax\ifmmode\mathchoice
8: {\hbox{\cmss Z\kern-.4em Z}}{\hbox{\cmss Z\kern-.4em Z}}
9: {\lower.9pt\hbox{\cmsss Z\kern-.4em Z}} {\lower1.2pt\hbox{\cmsss
10: Z\kern-.4em Z}}\else{\cmss Z\kern-.4em Z}\fi}
11: \def\IR{\relax{\rm I\kern-.18em R}}
12: \def\eg{{\it e.g.}}
13: \def\one{{\hbox{ 1\kern-.8mm l}}}
14: \def\gh{{\rm gh}}
15: \def\sgh{{\rm sgh}}
16: \def\NS{{\rm NS}}
17: \def\R{{\rm R}}
18: \def\ii{{\rm i}}
19: \def\bz{{\bar z}}
20: \def\comm#1#2{\left[ #1, #2\right]}
21: \def\acomm#1#2{\left\{ #1, #2\right\}}
22: \def\tr{{\rm tr\,}}
23: \newcommand{\N}{{\cal N}}
24: \newlength{\bredde}
25: \def\slash#1{\settowidth{\bredde}{$#1$}\ifmmode\,\raisebox{.15ex}{/}
26: \hspace*{-\bredde} #1\else$\,\raisebox{.15ex}{/}\hspace*{-\bredde}
27: #1$\fi}
28: \newcommand{\ft}[2]{{\textstyle\frac{#1}{#2}}}
29: \newcommand  {\Rbar} {{\mbox{\rm$\mbox{I}\!\mbox{R}$}}}
30: \newcommand  {\Hbar} {{\mbox{\rm$\mbox{I}\!\mbox{H}$}}}
31: \newcommand {\Cbar}
32:     {\mathord{\setlength{\unitlength}{1em}
33:      \begin{picture}(0.6,0.7)(-0.1,0)
34:         \put(-0.1,0){\rm C}
35:         \thicklines
36:         \put(0.2,0.05){\line(0,1){0.55}}
37:      \end {picture}}}
38: \newsavebox{\zzzbar}
39: \sbox{\zzzbar}
40:   {\setlength{\unitlength}{0.9em}
41:   \begin{picture}(0.6,0.7)
42:   \thinlines
43:   \put(0,0){\line(1,0){0.6}}
44:   \put(0,0.75){\line(1,0){0.575}}
45:   \multiput(0,0)(0.0125,0.025){30}{\rule{0.3pt}{0.3pt}}
46:   \multiput(0.2,0)(0.0125,0.025){30}{\rule{0.3pt}{0.3pt}}
47:   \put(0,0.75){\line(0,-1){0.15}}
48:   \put(0.015,0.75){\line(0,-1){0.1}}
49:   \put(0.03,0.75){\line(0,-1){0.075}}
50:   \put(0.045,0.75){\line(0,-1){0.05}}
51:   \put(0.05,0.75){\line(0,-1){0.025}}
52:   \put(0.6,0){\line(0,1){0.15}}
53:   \put(0.585,0){\line(0,1){0.1}}
54:   \put(0.57,0){\line(0,1){0.075}}
55:   \put(0.555,0){\line(0,1){0.05}}
56:   \put(0.55,0){\line(0,1){0.025}}
57:   \end{picture}}
58: \newcommand{\Zbar}{\mathord{\!{\usebox{\zzzbar}}}}
59: \def\Im{{\rm Im ~}}
60: \def\Re{{\rm Re ~}}
61: \newcommand{\bra}[1]{\langle{#1}|}
62: \newcommand{\ket}[1]{|{#1}\rangle}
63: \newcommand{\braket}[2]{\langle{#1}|{#2}\rangle}
64: \newcommand{\ena}{\end{eqnarray}}
65: \newcommand{\beqa}{\begin{eqnarray}}
66: \newcommand{\eeqa}{\end{eqnarray}}
67: \newcommand{\sect}[1]{Section~\ref{#1}}
68: \newcommand{\shalf}{\frac{1}{2}}
69: \newcommand{\eq}[1]{(\ref{#1})}
70: \newcommand{\fig}[1]{Fig.~\ref{#1}}
71: \newcommand{\chap}[1]{Chapter~\ref{#1}}
72: \newcommand{\be}{\begin{equation}}
73: \newcommand{\ee}{\end{equation}}
74: \def\G{\Gamma}
75: \def\K{{\cal K}}
76: \def\N{{\cal N}}
77: \def\H{{\cal H}}
78: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
79: \begin{document}
80: 
81: \begin{titlepage}
82: \begin{flushright}
83: EFI-03-30 \\
84: %UP \\
85: hep-th/0308057
86: \end{flushright}
87: %\vfill
88: \begin{center}
89: {\LARGE\bf Global Fluctuation Spectra \\ \vskip 3mm in Big Crunch/Big Bang String Vacua}    \\
90: \vskip 10.mm
91: {  Ben Craps $^{1}$ and Burt A.~Ovrut $^2$ } \\
92: \vskip 7mm
93: {\em $^1$ Enrico Fermi Institute, University of Chicago, 5640 S.~Ellis Av., Chicago, IL 60637, USA}\\
94: \vskip 0.5cm
95: {\em $^2$ Department of Physics, University of Pennsylvania, Philadelphia, PA 19104--6396}
96: \end{center}
97: \vfill
98: 
99: \begin{center}
100: {\bf ABSTRACT}
101: \end{center}
102: %\begin{quote}
103: We study Big Crunch/Big Bang cosmologies that correspond to exact world-sheet
104: superconformal field theories of type II strings. The string
105: theory spacetime contains a Big Crunch and a Big Bang cosmology,
106: as well as additional ``whisker'' asymptotic and intermediate regions.
107: Within the context of free string theory, we compute,
108: unambiguously, the scalar fluctuation spectrum in all regions of spacetime. Generically,
109: the Big Crunch fluctuation spectrum is altered
110: while passing through the bounce singularity. The change in the spectrum is
111: characterized by a function $\Delta$, which is momentum and time-dependent.
112: We compute $\Delta$ explicitly and
113: demonstrate that it arises from the whisker regions.
114: The whiskers are also shown to lead to ``entanglement'' entropy in
115: the Big Bang region.
116: Finally, in the Milne orbifold limit of our superconformal vacua, we show that
117: $\Delta\rightarrow 1$ and, hence, the fluctuation spectrum is unaltered by the
118: Big Crunch/Big Bang singularity. We comment on, but do not attempt
119: to resolve, subtleties related to gravitational backreaction and
120: light winding modes when interactions are taken into account.
121: %\vskip 7mm
122: \vfill
123: \hrule width 5.cm
124: \vskip 2.mm
125: {\small
126: \noindent  {\tt craps@theory.uchicago.edu,  ovrut@ovrut.hep.upenn.edu}}
127: %\end{quote}
128: %\begin{flushleft}
129: %PACS 11.25.-w, 04.65.+e
130: %\end{flushleft}
131: \end{titlepage}
132: 
133: 
134: 
135: \section{Introduction}
136: 
137: The theory of Ekpyrotic cosmology~\cite{cosmoA,cosmoB,cosmoC}, as well as
138: pre Big Bang scenarios~\cite{Gasperini:2002bn}, has emphasized the idea that the Universe
139: did not begin at the Big Bang but, rather, had a long prior history. Although the
140: details of this theory are far from understood, it is not
141: unreasonable to assume, since the Universe had to pass through the Big Bang
142: where densities and temperatures are set by the Planck scale, that
143: superstrings play a fundamental role. If this is the case, then one expects
144: the natural setting for cosmology to be, not four dimensions, but
145: the higher-dimensional spacetime of string theory. Ekpyrotic cosmology
146: introduced the idea that the Big Bang perhaps resulted from the
147: catastrophic collision of two brane solitons in this higher dimensional
148: space. Colliding branes and the associated bulk space geometry correspond to
149: vacuum solutions of the string theory equations of motion. These vacua may
150: solve equations which are valid to some finite order in a string expansion parameter,
151: such as five-brane/nine-brane collisions in heterotic $M$-theory~\cite{cosmoD,cosmoE} which are
152: valid to order $\kappa_{11}^{2}$, or they may be
153: exact conformal field theories, which solve the string equations to
154: all orders in the string parameter $\alpha'$. The second type of vacua were
155: emphasized in the Big Crunch/Big Bang~\cite{Khoury:2001bz} realizations of Ekpyrotic theory. Cyclic
156: models~\cite{cosmoF} are based on Big Crunch/Big Bang theories. All of these vacua
157: have in common the property that, prior to the Big Bang, the relevant region
158: of spacetime, our past, is contracting toward a singular brane collision.
159: In Big Crunch/Big Bang scenarios, the Universe then expands outward from this
160: singularity as the Big Bang. The singular point is called the ``bounce''.
161: 
162: 
163: Of particular importance in all cosmologies, and no less so in Ekpyrotic
164: theories, is the origin and momentum spectra of both scalar field and
165: gravitational quantum fluctuations. These are of the utmost importance since
166: they produce inhomogeneities in the cosmic microwave background (CMB) that have
167: already been observed. These observations are being increasingly refined and
168: offer an experimental window into whatever fundamental physics is responsible
169: for the present Universe. It was shown in~\cite{cosmoA,cosmoC} that nearly scale invariant
170: scalar perturbations are generated in the Big Crunch phase of Ekpyrotic
171: theories prior to the Big Bang. Indeed, a near scale invariant spectrum of
172: perturbations is observed in the CMB, but this is in the post Big Bang phase.
173: To make contact with these observations, one one must conclusively demonstrate
174: that the pre Big Bang fluctuation spectrum in Ekpyrotic theories
175: is propagated, nearly unchanged, through the Big Bang. This, despite the fact that
176: Big Crunch/Big Bang geometries are singular at the bounce. Although motivated
177: by string theory, almost all previous attempts to address this issue have been
178: carried out within the context of four-dimensional toy model geometries,
179: which do not solve the string equations of motion, even to lowest order~\cite{cosmoC,no,maybe}.
180: Recently, this problem was studied in a five-dimensional Milne background~\cite{Tolley:2002cv,
181: Tolley:2003nx}, and it
182: was emphasized that the five-dimensional structure is important near the Crunch.
183: However, in all cases the problem was studied
184: using the techniques of quantum field theory, with no string theory input.
185: With no further information at their disposal, the authors of these papers had
186: to proceed by matching the relevant pre Big Bang and post Big Bang wavefunctions
187: at the bounce, where the geometry is highly singular. The choice of boundary
188: conditions is, by itself, conjectural, being motivated by differing physical
189: arguments. This is made all the harder by the singular nature of the geometry
190: at the bounce. The result is that some authors claimed that the pre Big Bang
191: scale invariant fluctuations are radically altered when they pass through
192: the singularity~\cite{no} while others claim they are unaltered~\cite{cosmoC,Tolley:2003nx}. The ambiguous
193: nature of these attempts was demonstrated in~\cite{maybe}, who showed that, although
194: highly restricted, alternative boundary conditions are possible, leading to
195: the contradictory claims in the literature. How, then, can one resolve this
196: ambiguity?
197: 
198: 
199: It seems clear from the previous discussion that string theory should lead to
200: a unique solution of this problem. The reason for this is that string theory
201: vacua are, in general, globally defined. First of all, their geometric
202: manifolds include both the Big Crunch and Big Bang regions of cosmology, in
203: addition to other regions. Secondly, the wavefunctions of these vacua are
204: defined everywhere on the geometry. That is, knowing a wavefunction in the Big
205: Crunch region, for example, uniquely specifies the wavefunction in the
206: Big Bang regime. Clearly, this exactly specifies the boundary conditions for
207: the wavefunctions at the singularity, completely resolving the ambiguities
208: present in previous work. In this paper, we will show that this is indeed the
209: case, at least in the limit of zero string coupling. Working within the
210: framework of both supercritical \cite{Myers:fv,Antoniadis:1988vi} and critical type~II
211: superstrings\footnote{Note that ``type~II'' refers to a fermionic string with a chiral
212: GSO projection.}, we present a class of Big Crunch/Big Bang
213: cosmological vacua, called ``generalized'' Milne orbifolds \cite{Craps:2002ii}.
214: The geometry of these vacua includes both the past Big
215: Crunch and the future Big Bang regions. There are, in addition, four other
216: regions, often called ``whiskers'': an additional early time region, an additional
217: late time region and two intermediate regions with closed timelike curves. The
218: intermediate regions connect to the Big Crunch/Big Bang
219: regions at the bounce singularity. These vacua are exact superconformal field
220: theories and our results are valid to all
221: orders in the string worldsheet parameter $\alpha'$ \cite{Craps:2002ii}.
222: %\footnote{Or
223: %more precisely, our derivation of the fluctuation spectra is valid
224: %to leading order in $\alpha'$, but $\alpha'$ corrections do not
225: %change the results \cite{Craps:2002ii}.}
226: In this paper, we will present our discussion within the context of supercritical
227: string theory, since the corresponding spacetimes are manifestly homogeneous and
228: isotropic. However, we can show that the results are identical for
229: certain solutions of critical string theory. These ten-dimensional vacua contain, in
230: addition to the generalized Milne directions, two non-compact
231: spatial directions described by the
232: two-dimensional ``cigar'' conformal field theory \cite{Witten:1991yr},
233: which we take to be two of the three large
234: space directions. These vacua are not homogeneous and
235: isotropic, but approach these properties, as closely as one likes,
236: if a parameter is taken to be small. The Milne orbifold can be obtained as a specific limit
237: of these generalized Milne orbifolds. Working within this context, we find
238: that the wavefunctions are, indeed, globally defined. This follows from the
239: fact that the generalized Milne vacua all descend from a ``covering space'' by
240: the process of cosetting out a gauge action and orbifolding. The invariant globally
241: defined wavefunctions on the covering space then descend to globally defined
242: wavefunctions on the generalized Milne orbifolds. Therefore, to all orders in
243: $\alpha'$, the boundary condition ambiguities inherent in previous work have been
244: resolved.
245: 
246: What, then, are the results for the fluctuation spectrum? We find that the
247: fluctuation spectrum in the Big Crunch regime is, in general, altered by its
248: passage through the Big Crunch/Big Bang singularity. The change in the spectrum
249: can be calculated unambiguously at any time after the bounce. In the far future,
250: it can be expressed by an explicit momentum and time-dependent function
251: $\Delta(\vec{k},t)$,
252: which multiplies the early time pre Big Bang fluctuation spectrum. In the
253: Milne orbifold limit, we find that $\Delta\rightarrow 1$ and, hence, the
254: Big Crunch fluctuation spectrum is preserved as it passes through the
255: singularity. This proves the conjecture first introduced in~\cite{cosmoC} and shown
256: within a five-dimensional field theory context in~\cite{Tolley:2003nx}. If this result survives
257: corrections due to gravitational backreaction and stringy effects, which we will
258: comment on at the end of this Introduction, it means that the fluctuation
259: spectrum of Ekpyrotic cosmology may well be consistent with observation.
260: However, for generalized Milne orbifolds $\Delta$ is not unity and the
261: fluctuation spectrum changes as it passes through the bounce singularity. We
262: will show that this change is entirely due to the existence of
263: the whisker regions. Specifically, the quantum mechanics of these stringy
264: regions is inextricably linked to the quantum mechanics of the Big Crunch/Big
265: Bang geometry. The point of linkage is at the bounce singularity, where the
266: whiskers and the Big Crunch/Big Bang regions touch. The result is that
267: there is, in general, particle production in the future Big Bang region, even
268: though the vacuum in the Big Crunch was empty. This particle creation affects
269: all of the correlation functions in the future. In particular, it affects the
270: two-point correlation function from which $\Delta$ is calculated. We find that
271: $\Delta$ is an explicit function of momentum, with both time-independent and
272: time-dependent components. This calculation exposes a subtlety that arises
273: whenever the geometry has whiskers of the type discussed in this paper. That
274: is, that the early time in-vacuum is ambiguous. If we define this vacuum to be
275: such that an observer in the Big Crunch regime sees no particles, then we find
276: the associated in-vacuum in the early time whisker is not completely fixed. This
277: ambiguity in the in-vacuum can be parameterized by two constrained complex functions
278: $\gamma(\vec k)$ and $\tilde\gamma(\vec k)$. The parameter $\gamma(\vec k)$
279: explicitly enters the expression for $\Delta$. One natural choice turns out
280: to correspond to $\gamma(\vec k)=1, \tilde\gamma(\vec k)=0$ and leads to particle
281: creation and a
282: deviation of $\Delta$ from unity. However, as $\gamma\rightarrow 0$, the
283: particle production in the Big Bang region decreases to zero and $\Delta\rightarrow 1$. Therefore,
284: for a large choice of in-vacua, the change in the fluctuation spectrum by the
285: singularity is small. This might open the possibility of measuring string effects,
286: namely the existence of whisker regions connected to the Big
287: Crunch/Big Bang geometry, as small momentum and time-dependent deviations
288: from the scale invariance of the inhomogeneities in the CMB. We
289: hope to discuss this further elsewhere.
290: 
291: 
292: The linkage of the quantum mechanics of the Big
293: Crunch/Big Bang and whisker regions has a second, unanticipated effect.
294: As viewed in the Hilbert space of the complete quantum mechanics, the in-vacuum
295: is a pure state. However, we find that it is ``entangled'', that is, contains
296: correlations between the late time Big Bang and whisker regions. Therefore,
297: tracing its density matrix over the states of the unobserved whisker,
298: one obtains a non-trivial density matrix in the observable Big Bang region.
299: That is, an observer in the Big Bang region, with no access to information
300: about the whiskers, finds himself in a mixed state. This state has
301: ``entanglement''entropy, which manifests itself explicitly in the expression
302: for $\Delta$. If the entropy were zero, then the expression for $\Delta$
303: should be compatible with a Bogolubov transformation linking the ``in'' and
304: ``out'' states of the Big Crunch and Big Bang regions. However, non-vanishing
305: entanglement entropy will ruin this compatibility. We find that our explicit
306: expression for $\Delta$ indicates non-vanishing entanglement entropy, except
307: in the limit $\gamma\rightarrow 0$ where the whisker effects decouple. It
308: follows that, not only are the changes in the spectrum
309: calculable, but they explicitly exhibit entropy induced by the
310: existence of the stringy whisker regions. This is not dissimilar to recent
311: discussions of entangled states within the context of BTZ black holes
312: \cite{Maldacena:2001kr}, see also \cite{BTZ}.
313: 
314: We have presented the results of this paper mostly in the context
315: of Ekpyrotic Big Crunch/Big Bang transitions. Indeed, Ekpyrotic
316: theory has inspired much of the recent interest in Big Crunch/Big Bang
317: singularities in string theory \cite{Khoury:2001bz,Seiberg:2002hr}
318: and presents a context where it is particularly
319: important to know what happens to fluctuations at such a singularity.
320: However, Big Crunch/Big Bang transitions are of more general
321: interest and are especially appealing in string theory, where
322: observables are usually defined using simple asymptotic regions.
323: See, for instance, \cite{Gasperini:2002bn}.
324: Therefore, we believe that the results of this paper, such as
325: information loss due to the existence of  whisker regions, might be of interest in
326: more general cosmological scenarios.
327: 
328: 
329: Specifically, in this paper we do the following. Section 2 is devoted to a
330: discussion of a specific four-dimensional quantum field theory involving
331: gravity and three scalar fields. In subsection 2.1, we present a cosmological
332: background solution of this theory and explore various physical properties.
333: It is shown that there are two independent branches to
334: this solution, one representing a Big Crunch universe, evolving from the past
335: and ending in a singularity, and the other a Big Bang universe, beginning in a
336: singularity and evolving into the future. Scalar field quantum fluctuations, in
337: each of the Big Crunch and Big Bang regions, are discussed in subsection 2.2.
338: Both the in-vacuum and the out-vacuum are defined, and we explicitly compute
339: the scalar two-point correlation function with respect to each of
340: these vacua. In subsection 2.3, we relate both the classical and quantum
341: theories of the Big Crunch and Big Bang regions by connecting them at the
342: singularity. Within this context, we compute the two-point correlation
343: function with respect to the in-vacuum in both regions and compare them. We
344: find that the scalar fluctuation spectrum is potentially altered when it
345: passes from the past Big Crunch region through the bounce singularity. In the
346: far future, the change in the spectrum can be expressed by a momentum and
347: time-dependent function. We compute this function explicitly and show that it
348: depends on the Bogolubov coefficients relating the ``in'' and the ``out''
349: states of the Hilbert space. The meaning of these coefficients, and how one
350: should compute them, is discussed. Finally, in subsection 2.4, we show that
351: our four-dimensional quantum field theory is the low energy effective theory
352: for supercritical type II strings.
353: 
354: 
355: Section 3 is devoted to superstring cosmology within the context of type II
356: supercritical string theories. In subsection 3.1, we review the
357: ``generalized'' Milne orbifold solutions of these theories first presented in
358: \cite{Craps:2002ii}.
359: These solutions are exact superconformal field theories. The various
360: regions of the associated spacetimes are discussed in detail, including the
361: additional whisker regions. It is shown that, at low energy,
362: these string vacua give rise to the Big Crunch/Big Bang theories introduced in
363: section 2. We study the scalar fluctuation spectrum in subsection 3.2. The
364: structure of generalized Milne orbifolds as the coset and orbifold of a
365: $PSL(2,\Rbar)$ covering space is reviewed and the relationship of these vacua
366: to the Milne orbifold is discussed. In subsection 3.2.1, a basis of
367: wavefunctions, each defined on every region of the spacetime, is presented and
368: studied. We quantize the scalar fluctuations in subsection 3.2.2. This is
369: accomplished by expanding the scalar fluctuations in this basis and
370: canonically quantizing the coefficients. We define two natural vacua,
371: the in-vacuum and the out-vacuum, and compute the Bogolubov coefficients
372: relating them \cite{Craps:2002ii}. Using these results, we compute the particle production in the
373: future regions and the scalar two-point correlation function with respect to
374: the in-vacuum. This result is valid in any region, allowing us to compare the
375: spectrum in the future Big Bang region with the early time spectrum
376: during the Big Crunch. We find that, generically, the spectrum is altered. In
377: the far future, we can express the change in the spectrum in terms of a
378: function $\Delta$, which is momentum and time-dependent. We compute this
379: function explicitly. In subsection 3.2.3, we show that there is, in fact, a
380: family of in-vacua, which we specify with two constrained complex functions
381: $\gamma(\vec k)$ and $\tilde\gamma(\vec k)$. We repeat the calculation of the Bogolubov coefficients, particle
382: production, scalar two-point correlation function and $\Delta$ in this context.
383: Again, in general, the fluctuation spectrum is changed as it passes through
384: the bounce singularity. However, in the limit that $\gamma \rightarrow 0$,
385: $\Delta$ approaches unity and the spectrum is conserved. In this limit,
386: $\Delta$ can be expressed in terms of the pure Big Crunch/Big Bang Bogolubov
387: coefficients. However, for any finite $\gamma$, this is no longer the case.
388: This is explained in subsection~3.3, where it is shown that the generalized
389: in-vacuum is an entangled state in the future. This implies that, from the point
390: of view of an observer in the Big Bang region, this vacuum has entanglement
391: entropy. This is equivalent to quantum mechanical ``information loss'' into
392: the whisker regions. It is this information loss that obstructs writing
393: $\Delta$ in terms of the pure Big Crunch/Big Bang Bogolubov coefficients. In
394: subsection 3.4, we discuss the limit of our superconformal vacua to the Milne
395: orbifold. It is shown that, in this limit, the factor $\Delta$ goes to unity
396: for any $\gamma$ in-vacuum. This proves that, in the Milne orbifold
397: cosmology, the fluctuation spectrum is unaltered by the bounce
398: singularity. In subsection~3.5, we include some comments on backreaction.
399: Finally, in Appendices A, B, and C, we outline the theory of
400: global wavefunctions, present the expressions for a basis of these
401: wavefunctions in all regions of the spacetime and discuss the Milne limit of
402: these wavefunctions respectively.
403: 
404: 
405: Before proceeding, we would like to make several important comments. First,
406: note that the generalized Milne orbifolds
407: have a smooth circle in the spatial Milne-direction.
408: However, it is not hard to show that all of the results of this paper will be
409: unchanged if we allow a further $\Zbar_{2}$ orbifolding in this direction. In
410: this case, this smooth circle becomes a finite interval, with new twisted
411: sector states appearing on each of the two boundaries. This new $\Zbar_{2}$
412: orbifolded vacuum is again an exact superconformal field theory, which is
413: closer in spirit to the notion of colliding branes.
414: 
415: As we have stated previously, our results are computed with respect to  exact
416: superconformal field theories, so $\alpha'$ corrections are under control.
417: However, it is well-known that these theories suffer from severe backreaction
418: problems, which could conceivably modify our results.
419: Indeed, string perturbation theory gives rise not only
420: to an expansion in $\alpha'$ (higher derivative corrections to the classical action),
421: but also to an expansion in the string coupling $g_s$
422: (quantum corrections). The results of \cite{Craps:2002ii, Elitzur:2002rt} on
423: global wavefunctions refer only to the first term in the $g_s$ expansion. It has,
424: in fact, been shown in \cite{Berkooz:2002je}, following \cite{Liu:2002ft}, that classical
425: string scattering amplitudes exhibit divergences associated with a large backreaction of
426: the spacetime geometry to small perturbations. For a non-perturbative manifestation of
427: gravitational instability, see \cite{Horowitz:2002mw}.
428: Large gravitational backreaction
429: (or the absence thereof) in this and related models \cite{Elitzur:2002rt, Cornalba:2002fi,
430: related}
431: was also studied in \cite{backreaction}.
432: The tree-level divergences of \cite{Berkooz:2002je, Liu:2002ft} indicate a breakdown of
433: string perturbation theory. That is, it is not consistent to
434: ignore $g_s$ corrections if $g_s$ is small but non-zero. For a non-technical discussion
435: of this, see \cite{Liu:2002yd}.
436: Therefore, there is at present no fully controlled computational framework, and it is
437: conceivable that our detailed results will receive significant corrections when
438: backreaction is taken into account. One may hope that at least certain important
439: qualitative features, such as the peculiar causal structure of the string solutions
440: and its effect on four-dimensional cosmology, will survive $g_s$ corrections. However, it
441: will take significant advances in string theory to either establish or refute this.
442: 
443: It has been suggested, see for instance \cite{Bachas:2002qt}, that string winding
444: modes that become light
445: near the bounce might play a role in resolving the singularity in
446: string theory. In the case of the Milne orbifold, the description
447: of these winding modes is somewhat complicated, because the size of the
448: Milne circle grows without bound away from the singularity.
449: See \cite{Berkooz:2003bs} for a very recent discussion and
450: \cite{Nekrasov:2002kf} for earlier work. However, in the generalized Milne
451: orbifold the radius of the Milne circle approaches a constant
452: asymptotically, and the vertex operators for the winding modes are
453: explicitly known \cite{Craps:2002ii}. In the present paper, we
454: will ignore winding modes, except to note that they become light
455: near the Big Crunch/Big Bang singularity and invalidate a four-dimensional description
456: there. They are heavy away from the bounce and, thus, do not appear in the
457: effective action describing our four-dimensional cosmology. However, we
458: should stress that the computations we do for the generalized Milne orbifold
459: can easily be extended to include winding modes
460: \cite{Craps:2002ii}. Of course, the most interesting effects of winding modes
461: should involve string interactions, and these have not yet been
462: computed for the generalized Milne orbifold.
463: 
464: 
465: In string theory, the observables are S-matrix elements. At zero string coupling, the
466: regime
467: in which we will be working, they are determined by a Bogolubov
468: matrix. Our strategy is to compute this matrix using the globally defined vertex operators
469: \cite{Craps:2002ii, Elitzur:2002rt}. More specifically, we calculate the Bogolubov
470: matrix to leading order in $\alpha'$ from the associated globally defined
471: wavefunctions. However, as was mentioned in \cite{Craps:2002ii},
472: the higher $\alpha'$ corrections modify this matrix in a very
473: trivial way, simply multiplying some of the entries by a phase.
474: Therefore, the particle creation rates we compute are exact to all
475: orders in $\alpha'$. In this paper, we will go a step beyond computing
476: S-matrix elements and compute correlation functions at a fixed finite time,
477: using the global wavefunctions we obtained from string theory.
478: 
479: %On the other hand, finite-time correlation functions, the objects of
480: %interest from an experimental point of view, are hard to define, let alone compute,
481: %in string theory: they are believed to be derived concepts, encoded in a very
482: %complicated way in the fundamental observables of the theory (S-matrix-type quantities).
483: %Therefore, when we compute fluctuation spectra at some finite time, we strictly speaking
484: %leave the rigorous framework of string theory, and use the
485: %string theory results we derived, the global wavefunctions, as an ingredient in
486: %field theory computations. This, however, is a complication that
487: %is not associated to the Big Crunch/Big Bang singularity: it would
488: %also be there for computations in Minkowski space.
489: 
490: For additional recent string theory work related to Big Crunch/Big Bang singularities,
491: see \cite{recent}.
492: 
493: 
494: 
495: 
496: \section{Four-Dimensional Cosmology}\label{sec:fourd}
497: 
498: 
499: In this section, we will explore the cosmological properties of a specific
500: four-dimensional quantum field theory coupled to gravity.
501: This theory consists of three scalar
502: fields, denoted by $\sigma_{T}$, $\sigma_{R}$ and $\phi$ respectively, coupled to
503: the usual Einstein gravity. The action for this theory is given by
504: \begin{equation}
505: S=\frac{1}{2\kappa_{4}^{2}}\int d^{4}x\sqrt{-g}{\cal{L}},
506: \label{eq:1}
507: \end{equation}
508: where
509: \begin{equation}
510: {\cal{L}}=R-\frac{1}{2}g^{\mu\nu}\partial_{\mu}\phi\partial_{\nu}\phi
511: -\frac{1}{2}g^{\mu\nu}\partial_{\mu}\sigma_{T}\partial_{\nu}\sigma_{T}
512: -\frac{1}{2}g^{\mu\nu}\partial_{\mu}\sigma_{R}\partial_{\nu}\sigma_{R}
513: -4Q^2e^{\phi}.
514: \label{eq:2}
515: \end{equation}
516: There are two dimensionful parameters in this action, Newton's constant
517: $\kappa_{4}^{2}/8\pi$ ($\kappa_4$ having the dimension of a length)
518: and a mass scale $Q$, which we will think of as small compared to the
519: momenta of interest.
520: 
521: 
522: \subsection{The Background Spacetime}\label{sub:background}
523: 
524: We will be interested in cosmological solutions of the equations of motion
525: of this theory that are spatially homogeneous and isotropic, that is, of the
526: Friedman-Robinson-Walker (FRW) type. We find the following general class of such
527: solutions. The scalar fields are independent of all spatial coordinates and
528: evolve as follows. The field $\sigma_{T}$ simply vanishes,
529: \begin{equation}
530: \sigma_{T}=0,
531: \label{eq:3}
532: \end{equation}
533: whereas
534: \begin{equation}
535: \sigma_{R}=\sqrt2\log|\tanh(Qt)|
536: \label{eq:4}
537: \end{equation}
538: and
539: \begin{equation}
540: \phi=2\phi_{0}-\log|\sinh(2Qt)|.
541: \label{eq:5}
542: \end{equation}
543: Here
544: %\begin{equation}
545: %Q=\sqrt{\frac{\Lambda}{2}}
546: %\label{eq:6}
547: %\end{equation}
548: %and
549: $\phi_{0}$ is an arbitrary integration constant. These fields drive a
550: time-varying metric which is independent of all spatial coordinates and given
551: by
552: \begin{equation}
553: g_{\mu\nu}=a(t)^{2}\eta_{\mu\nu},
554: \label{eq:7}
555: \end{equation}
556: where
557: \begin{equation}
558: a^{2}=e^{-2\phi_{0}}|\sinh(2Qt)|
559: \label{eq:8}
560: \end{equation}
561: and $\eta_{\mu\nu}$ is the metric of flat Minkowski space. It follows from the
562: form of the metric that $t$ is conformal time.
563: 
564: 
565: What are the cosmological properties of this class of solutions? Perhaps the
566: most salient feature is that the conformal factor of the metric, $a(t)^{2}$,
567: is monotonically decreasing in the time interval $-\infty < t < 0$ and
568: monotonically increasing for $0<t< +\infty$. We will refer to the spacetime in the
569: negative time interval as Region I and the spacetime for the positive time
570: interval as Region II. Let us be more specific about the geometry of these
571: regions. To make contact with the conventional analysis of FRW cosmologies, we
572: will first compute the equation of state
573: \begin{equation}
574: w=\frac{P}{\rho},
575: \label{eq:9}
576: \end{equation}
577: where $P$ is the pressure and $\rho$ the energy density. These quantities are
578: defined by the Einstein equations
579: \begin{equation}
580: {\cal{H}}^{2}=\frac{\kappa_{4}^{2}}{3}a^{2}\rho
581: \label{eq:10}
582: \end{equation}
583: and
584: \begin{equation}
585: {\cal{H}}'=-\frac{\kappa_{4}^{2}}{6}a^{2}(\rho+3P),
586: \label{eq:11}
587: \end{equation}
588: where
589: \begin{equation}
590: {\cal{H}}=\frac{a'}{a}
591: \label{eq:12}
592: \end{equation}
593: and the prime denotes differentiation with respect to conformal time $t$.
594: Using~\eq{eq:8}, we find from~\eq{eq:10} and~\eq{eq:11} that
595: \begin{equation}
596: \rho= \frac{3Q^{2}}{\kappa_{4}^{2}}e^{2\phi_{0}}\frac{\cosh^{2}(2Qt)}{|\sinh^{3}
597: (2Qt)|}
598: \label{eq:13}
599: \end{equation}
600: and
601: \begin{equation}
602: P=\frac{Q^{2}}{\kappa_{4}^{2}}e^{2\phi_{0}}\left(\frac{4-\cosh^{2}(2Qt)}{|\sinh^{3}
603: (2Qt)|}\right)
604: \label{eq:14}
605: \end{equation}
606: respectively. It follows that
607: \begin{equation}
608: w=\frac{1}{3}\left(\frac{4-\cosh^{2}(2Qt)}{\cosh^{2}(2Qt)}\right).
609: \label{eq:15}
610: \end{equation}
611: The first thing to notice is that $w$ is not constant in time. In the far past
612: and future
613: \begin{equation}
614: w\longrightarrow -\frac{1}{3}, \qquad t\longrightarrow \pm\infty.
615: \label{eq:16}
616: \end{equation}
617: As the cosmology evolves, one finds that
618: \begin{equation}
619: w=0, \qquad t_{0}=\pm\frac{1}{2Q}\log(2+\sqrt{3})\Leftrightarrow \cosh(2Qt_{0})=2
620: \label{eq:17}
621: \end{equation}
622: and
623: \begin{equation}
624: w=\frac{1}{3}, \qquad
625: t_{1/3}=\pm\frac{1}{2Q}\log(1+\sqrt{2})\Leftrightarrow
626: |\sinh(2Qt_{1/3})|=1.
627: \label{eq:18}
628: \end{equation}
629: Finally, as one approaches the origin, either from negative time or positive
630: time,
631: \begin{equation}
632: w\longrightarrow 1-\frac{16Q^{2}}{3}t^{2}+\cdots, \qquad t\longrightarrow \pm 0.
633: \label{eq:19}
634: \end{equation}
635: It follows that
636: \begin{equation}
637: w\cong 1, \qquad |t|\ll \frac{\sqrt{3}}{4Q}.
638: \label{eq:20}
639: \end{equation}
640: What is the interpretation of this somewhat unusual FRW cosmology? To
641: elucidate this, note that in a scalar dominated phase the energy density and
642: pressure are given by
643: \begin{equation}
644: \rho={1\over2\kappa_4^2}\left(\sum_{i=1}^{3}\frac{{\psi'_{i}}^{2}}{2a^2}+V\right),
645: \qquad
646: P={1\over2\kappa_4^2}\left(\sum_{i=1}^{3}\frac{{\psi'_{i}}^{2}}{2a^2}-V\right),
647: \label{eq:21}
648: \end{equation}
649: where $\psi_{i}, i=1,2,3$ are the fields $\sigma_{T}$, $\sigma_{R}$ and $\phi$
650: respectively and $V$ is the total potential energy.
651: Using~\eq{eq:2},~\eq{eq:3},~\eq{eq:4} and~\eq{eq:5},
652: it is easy to show that $\rho$ and $P$ in~\eq{eq:21} are identical to
653: expressions~\eq{eq:13} and~\eq{eq:14}. It follows that in both Regions I and
654: II our cosmology is
655: scalar dominated with the equation of state $w$ changing as the scalar fields
656: evolve. In the conformal time coordinate $t$, only in the regions described
657: by~\eq{eq:16} and~\eq{eq:20} does $w$ behave approximately as a constant, $w=-1/3$
658: and $w=1$ respectively.
659: Note that, in the latter time region, the conformal factor $a(t)$
660: in~\eq{eq:8} becomes
661: \begin{equation}
662: a\cong |\frac{t}{e^{2\phi_{0}}/2Q}|^{q}, \qquad q=\frac{1}{2},
663: \label{eq:22}
664: \end{equation}
665: which is consistent with constant $w=1$.
666: 
667: 
668: A second important feature of FRW cosmologies is the Hubble parameter
669: \begin{equation}
670: H=\frac{a'}{a^{2}}
671: \label{eq:23}
672: \end{equation}
673: and its inverse, $R_{H}=|H|^{-1}$, the Hubble radius. Using~\eq{eq:8}, one
674: finds that
675: \begin{equation}
676: H=Qe^{\phi_{0}}\frac{\cosh(2Qt)}{\sinh(2Qt)|\sinh(2Qt)|^{\frac{1}{2}}}.
677: \label{eq:24}
678: \end{equation}
679: Again, note that $H$ and, hence, $R_{H}$ are not constants but, rather,
680: evolve with conformal time $t$. Let us focus on the behaviour of the Hubble
681: radius. In the far past and future
682: \begin{equation}
683: R_{H}\longrightarrow \infty, \qquad t\longrightarrow \pm \infty.
684: \label{eq:25}
685: \end{equation}
686: The Hubble radius then changes monotonically, taking the value, for example,
687: \begin{equation}
688: R_{H}=\frac{e^{-\phi_{0}}}{\sqrt2Q}, \qquad t_{1/3}=\pm\frac{1}{2Q}
689: \log(1+\sqrt{2})
690: \label{eq:26}
691: \end{equation}
692: before behaving as
693: \begin{equation}
694: R_{H}\longrightarrow {|2Qt|^{3/2}e^{-\phi_0}\over Q}(1-Q^{2}t^{2}+\cdots),
695: \qquad t\longrightarrow \pm0
696: \label{eq:27}
697: \end{equation}
698: as the time approaches the origin. Hence,
699: \begin{equation}
700: R_{H}\cong {|2Qt|^{3/2}e^{-\phi_0}\over Q}, \qquad |t|\ll\frac{1}{Q},
701: \label{eq:28}
702: \end{equation}
703: which vanishes at $t=0$. Note that this behaviour is consistent with the
704: monotonic evolution of the conformal factor $a^{2}$ and is valid in both Region
705: I and II. From \eq{eq:8} and \eq{eq:28} it is clear that the Hubble radius goes to zero
706: more quickly than the physical wavelength of an excitation, which is proportional to $a$.
707: Therefore, all modes are ``frozen'' outside of the horizon at the Big Crunch/Big Bang transition.
708: Similarly, we conclude from \eq{eq:8} and \eq{eq:24} that at very early and very late times, modes
709: are inside or outside the horizon depending on whether their comoving momentum $\vec k$ satisfies
710: $\vec k^2>Q^2$ or $\vec k^2<Q^2$, respectively. In the following, we will focus on fluctuations with
711: $\vec k^2\gg Q^2$. Thus, these modes start out inside the horizon at very early times and freeze as they
712: approach the Big Crunch. Similarly, in the Big Bang region of spacetime they start out frozen
713: near the Big Bang and enter the horizon at some later time.
714: 
715: 
716: 
717: A third important quantity to consider is the scalar curvature, $R$.
718: In conformal time, the scalar curvature in FRW spacetimes
719: is given by
720: \begin{equation}
721: R=\frac{6a''}{a^{3}}.
722: \label{eq:29}
723: \end{equation}
724: It then follows from~\eq{eq:8} that
725: \begin{equation}
726: R=3Q^{2}e^{2\phi_{0}}\frac{\sinh^{2}(2Qt)-1}{|\sinh^{3}(2Qt)|},
727: \label{eq:30}
728: \end{equation}
729: an expression that is valid in both Regions I and II. Key values of the scalar
730: curvature occur at precisely the same times, namely  $\pm\infty$, $t_{0}$,
731: $t_{1/3}$ and $\pm0$ defined in~\eq{eq:17} and~\eq{eq:18} respectively,
732: that indicated the behaviour of the equation of state
733: $w$. In the far past and future
734: \begin{equation}
735: R \longrightarrow 0, \qquad t\longrightarrow \pm\infty.
736: \label{eq:31}
737: \end{equation}
738: As $|t|$ decreases from infinity toward zero, $R$ has the following
739: properties. To begin with, $R$ is positive
740: and increasing until
741: $|t_{0}|=\frac{1}{2Q}\log(2+\sqrt{3})$, after which $R$ monotonically decreases.
742: The scalar curvature vanishes at
743: \begin{equation}
744: R=0, \qquad  t_{1/3}=\pm\frac{1}{2Q}\log(1+\sqrt{2}).
745: \label{eq:32}
746: \end{equation}
747: For smaller values of $|t|$, $R$ is negative. As one approaches the origin,
748: \begin{equation}
749: R\longrightarrow
750: -\frac{3}{4}e^{2\phi_{0}}\frac{1}{|t|^{3}}(1-6Q^{2}t^{2}+\cdots), \qquad
751: t\longrightarrow \pm0
752: \label{eq:33}
753: \end{equation}
754: which diverges as
755: \begin{equation}
756: R\cong -\frac{3}{4}e^{2\phi_{0}}\frac{1}{|t|^{3}}, \qquad |t|\ll
757: \frac{1}{\sqrt{6}Q}.
758: \label{eq:34}
759: \end{equation}
760: 
761: 
762: 
763: The above results allow us to give a concise description of our
764: cosmological solution. First consider Region I, corresponding to the negative
765: time interval $-\infty<t<0$. The associated geometry is that of a spatially
766: homogeneous and isotropic FRW spacetime. In the distant past, the manifold has
767: vanishing scalar curvature and a divergent Hubble radius. As time progresses,
768: the scalar curvature first grows positively, reaches a maximum and then begins to
769: decrease, vanishing at a finite time $t_{1/3}$. Henceforth,
770: the curvature is negative, diverging as $t^{-3}$ as $t$ approaches the origin
771: $t=0$. Throughout the entire time interval $-\infty<t<0$, the Hubble radius is
772: monotonically shrinking from infinity to zero. Both the vanishing of the
773: Hubble radius and, particularly, the divergence of the scalar curvature as
774: $t \rightarrow 0$, tells us that Region I terminates abruptly at $t=0$.
775: Region I, therefore, is a classic example of what is called a ``Big Crunch''
776: cosmology. Region II, on the other hand, corresponding to the positive time
777: interval $0<t<+\infty$, is the exact mirror image of Region I
778: in the time direction. That is, Region II is a spatially homogeneous and
779: isotropic FRW spacetime that starts abruptly at $t=0$ with vanishing Hubble
780: radius and negatively infinite scalar curvature and then expands outward. The
781: scalar curvature increases from negative to zero to positive, reaches a
782: maximum and then decreases to zero as $t \rightarrow +\infty$. During the
783: entire time interval $0<t<+\infty$, the Hubble radius monotonically increases
784: from zero to infinity. Region II, therefore, is a classic example of a ``Big
785: Bang'' cosmology. It is essential to note that in general relativity,
786: because of the curvature singularity at $t=0$, there is no relationship between
787: Region I and Region II, each representing an independent cosmology.
788: 
789: 
790: 
791: \subsection{Fluctuations}\label{sub:fluctuations}
792: 
793: We now turn to a discussion of quantum fluctuations of the scalar fields
794: on the Big Bang/Big Crunch geometries presented above. To do this, we must first
795: expand $\sigma_{T}$, $\sigma_{R}$ and $\phi$ around their classical values,
796: which we now denote as $\langle  \sigma_{T}\rangle $,
797: $\langle  \sigma_{R}\rangle $ and $\langle  \phi\rangle $, given
798: in~\eq{eq:3},~\eq{eq:4} and~\eq{eq:5} respectively. That is,
799: \begin{equation}
800: \sigma_{T}=\langle  \sigma_{T}\rangle +\delta\sigma_{T}, \qquad
801: \sigma_{R}=\langle  \sigma_{R}\rangle +\delta\sigma_{R}, \qquad
802: \phi=\langle  \phi\rangle +\delta\phi.
803: \label{eq:35}
804: \end{equation}
805: Inserting each of these fields into its equation of motion, using
806: expressions~\eq{eq:7} and~\eq{eq:8} for the background metric and assuming
807: the fluctuations are of the form
808: \begin{equation}
809: \delta\sigma_{T}=\delta T(t) e^{i{\vec{k}}\cdot{\vec{x}}}, \qquad
810: \delta\sigma_{R}=\delta R(t) e^{ i{\vec{k}}\cdot{\vec{x}}}, \qquad
811: \delta\phi=\delta \Phi(t) e^{ i{\vec{k}}\cdot{\vec{x}}},
812: \label{eq:36}
813: \end{equation}
814: we find that
815: the fluctuation $\delta T$ is a solution of
816: \begin{equation}
817: \delta T''+2Q\coth(2Qt)\delta T'+{\vec{k}}^{2}\delta T=0,
818: \label{eq:37}
819: \end{equation}
820: $\delta R$ solves the same equation
821: \begin{equation}
822: \delta R''+2Q\coth(2Qt)\delta R'+{\vec{k}}^{2}\delta R=0
823: \label{eq:38}
824: \end{equation}
825: whereas
826: \begin{equation}
827: \delta\Phi''+2Q\coth(2Qt)\delta\Phi'+({\vec{k}}^{2}+4Q^2)\delta\Phi, \qquad
828: \delta\Phi\ll 1.
829: \label{eq:39}
830: \end{equation}
831: Note that the $\delta \Phi$ fluctuations will satisfy the same equation as
832: $\delta T$ and $\delta R$ for momenta
833: \begin{equation}
834: {\vec{k}}^{2}\gg 4Q^2.
835: \label{eq:40}
836: \end{equation}
837: Henceforth, we will restrict our discussion to this momentum regime. Since, in
838: this case, all three fluctuations are specified by the same equation, we will
839: simply focus on one of them, which we choose to be $\delta T$.
840: 
841: 
842: 
843: Let us search for solutions of the $\delta T$ fluctuation
844: equation~\eq{eq:37}. To do this, we must first specify the region of spacetime
845: in which we want to work. Begin by considering Region I, with negative conformal
846: time in the interval $-\infty<t<0$. For very early times, \eq{eq:37}
847: simplifies to
848: \begin{equation}
849: \delta T''-2Q\delta T'+{\vec{k}}^{2}\delta T=0.
850: \label{eq:41}
851: \end{equation}
852: It is easy to see that this has plane wave solutions of the form
853: \begin{equation}
854: \delta T_{\vec{k}}^{\pm}= C_{\vec{k}}e^{Qt}e^{\mp iE_{\vec{k}}t},
855: \label{eq:42}
856: \end{equation}
857: where
858: \begin{equation}
859: E_{\vec{k}}=\sqrt{{\vec{k}}^{2}-Q^{2}}
860: \label{eq:43}
861: \end{equation}
862: is the energy associated with momentum $\vec{k}$ and $C_{\vec{k}}$ is
863: a normalization constant. Note that $E_{\vec{k}}$ is a positive real
864: number in the momentum regime~\eq{eq:40} in which we are working. The
865: normalization constant can be determined using the scalar product
866: \begin{equation}
867: (\phi_{1},\phi_{2})=-i\int_{\Sigma}{\phi_{1}{\stackrel{\leftrightarrow}
868: {\partial}}_\mu\phi_{2}^{*} \sqrt{g_{\Sigma}}d\Sigma^{\mu}}
869: \label{eq:44}
870: \end{equation}
871: where $\Sigma$ is a space-like three-surface, $\sqrt{g_{\Sigma}}$ is the
872: volume element on that surface,
873: \begin{equation}
874: d\Sigma^{\mu}={\delta^{\mu}}_{0}\sqrt{-g^{00}}d{\vec{x}}
875: \label{eq:45}
876: \end{equation}
877: and $\phi_{i}$, $i=1,2$ are any two scalar functions. Using the metric given
878: in~\eq{eq:7} and~\eq{eq:8}, expression~\eq{eq:42} for $\delta T$
879: and~\eq{eq:44}, we find that
880: \begin{equation}
881: (\delta T_ {\vec{k}}^{\pm}e^{i\vec k\cdot x}, \delta
882: T_{\vec{k'}}^{\pm}e^{i\vec k'\cdot \vec x} )=\pm(2\pi)^3\delta(\vec k-\vec k')
883: |C_{\vec{k}}|^{2}
884: e^{-2\phi_{0}}E_{\vec{k}}. \label{eq:46}
885: \end{equation}
886: %where $V_{\Sigma}$ is the volume of the three-surface $\Sigma$.
887: Note that the Klein-Gordon norm obtained from~\eq{eq:44} can be either positive or negative
888: depending on the frequency of the scalar function.
889: Setting the right hand side of~\eq{eq:46} equal to $\pm(2\pi)^3\delta(\vec k-\vec k')$,
890: it follows that the normalization constant, up to a phase, is given by
891: \begin{equation}
892: C_{\vec{k}}=\frac{e^{\phi_{0}}}{\sqrt{E_{\vec{k}}}}.
893: \label{eq:47}
894: \end{equation}
895: Therefore, the normalized asymptotic plane wave solutions of~\eq{eq:41}
896: are given by
897: \begin{equation}
898: \delta
899: T_{\vec{k}}^{\pm}=\frac{e^{\phi_{0}}}{\sqrt{E_{\vec{k}}}}e^{Qt}
900: e^{\mp iE_{\vec{k}}t}, \qquad t\longrightarrow -\infty.
901: \label{eq:48}
902: \end{equation}
903: One can, in fact, solve the $\delta T$ fluctuation
904: equation~\eq{eq:37} for any value of $t$ in Region I. The result is found to
905: be
906: \begin{equation}
907: \delta
908: T_{\vec{k}}^{+}= \frac{4^je^{\phi_{0}}}{\sqrt{E_{\vec{k}}}}
909: (-z)^{j}F(-j,-j;-2j;\frac{1}{z}),
910: \label{eq:49}
911: \end{equation}
912: where
913: \begin{equation}
914: j=-\frac{1}{2} + i\frac{E_{\vec{k}}}{2Q}, \qquad z=-\sinh^{2}(Qt)
915: \label{eq:50}
916: \end{equation}
917: and $F(a,b;c;x)$ is the hypergeometric function $_{2}F_{1}$ (see Appendix~B for
918: its definition and some properties). In addition,
919: the hermitian conjugate
920: \begin{equation}
921: \delta T_{\vec{k}}^{-}=\delta T_{\vec{k}}^{+*}
922: \label{eq:51}
923: \end{equation}
924: is an independent solution of~\eq{eq:37}.
925: Using the facts that
926: \begin{equation}
927: z \longrightarrow -\frac{e^{-2Qt}}{ 4}, \qquad
928: F(-j,-j;-2j;\frac{1}{z})\longrightarrow 1\ \ {\rm as}\ \  t\longrightarrow -\infty,
929: \label{eq:52}
930: \end{equation}
931: we see that~\eq{eq:49} and~\eq{eq:51} approach the plane waves solutions
932: \eq{eq:48} in the far past. One can show that~\eq{eq:49} and~\eq{eq:51}
933: diverge logarithmically at $t=0$.
934: 
935: 
936: 
937: Combining these results with the first
938: expression in~\eq{eq:36}, we see that any fluctuation $\delta\sigma_{T}$ in
939: Region I can be written as
940: \begin{equation}
941: {\delta\sigma_{T}}^{I}=\int\frac{d^3k}{(2\pi)^{3/2}}
942: ({a_{\vec{k}}}^{I}\delta T_{\vec{k}}^{+}(t)
943: e^{i\vec{k}\cdot\vec{x}}+{a_{\vec{k}}}^{I*}\delta
944: T_{\vec{k}}^{-}(t)e^{-i\vec{k}\cdot\vec{x}}),
945: \label{eq:53}
946: \end{equation}
947: where ${a_{\vec{k}}}^{I}$ are arbitrary complex coefficients.
948: %In writing this expansion, we have used the fact that
949: %\begin{equation}
950: %\frac{1}{V_{\Sigma}}\sum_{\vec{k}}=\int{\frac{d^3k}{(2\pi)^{3}}}.
951: %\label{eq:54}
952: %\end{equation}
953: %It follows that, henceforth, one should drop the $1/V_{\Sigma}$ factor in $\delta
954: %{T_{\vec{k}}}^{\pm}$ when using expansion~\eq{eq:53}.
955: Note that the first function in this expansion, $\delta
956: T_{\vec{k}}^{+}(t)e^{i\vec{k}\cdot\vec{x}}$, corresponds to a pure positive
957: frequency plane wave in the far past. Similarly, in this limit, $\delta
958: T_{\vec{k}}^{-}(t)e^{-i \vec{k}\cdot\vec{x}}$ becomes a negative frequency plane
959: wave.
960: 
961: 
962: Thus far, our discussion of the fluctuations ${\delta \sigma_{T}}^{I}$ has been
963: strictly classical. However, the theory can be easily quantized by demanding
964: that ${\delta \sigma_{T}}^{I}$ and, hence, the coefficients
965: ${a_{\vec{k}}}^{I}$ be operators in a Hilbert space. For simplicity,
966: we will continue to denote these operators by ${\delta
967: \sigma_{T}}^{I}$ and ${a_{\vec{k}}}^{I}$, suppressing the usual ``hat''
968: notation. The quantization will be
969: canonical if we assume that
970: \begin{equation}
971: [{a_{\vec{k}}}^{I}, {a_{\vec{k'}}}^{\dagger I}]=\delta^{3}(\vec{k}-\vec{k'}),
972: \qquad [{a_{\vec{k}}}^{I}, {a_{\vec{k'}}}^{I}]=[{a_{\vec{k}}}^{\dagger I},
973: {a_{\vec{k'}}}^{\dagger I}]=0.
974: \label{eq:55}
975: \end{equation}
976: The vacuum state of the quantum theory is then defined as the normalized state
977: $|0\rangle _{in}$ satisfying
978: \begin{equation}
979: {a_{\vec{k}}}^{I}|0\rangle _{in}=0
980: \label{eq:56}
981: \end{equation}
982: for all momenta $\vec{k}$. There are many objects that can now be discussed in
983: this context. In this paper, we will focus primarily on the two-point
984: correlation function
985: \be\label{eq:56bis}
986: _{in}\langle 0|{\delta \sigma_{T}}^{I}(t,\vec{x})
987: {\delta \sigma_{T}}^{I}(t,\vec{x}+\vec r)|0\rangle _{in}.
988: \ee
989: Using~\eq{eq:53},~\eq{eq:55} and~\eq{eq:56}, we find that this function is
990: independent of $\vec{x}$ and given by
991: \beqa
992: _{in}\langle 0|{\delta \sigma_{T}}^{I}(t,\vec{x})
993: {\delta \sigma_{T}}^{I}(t,\vec{x}+\vec r)|0\rangle _{in}&=&\int\frac{d^3k}{(2\pi)^{3}}
994: |\delta T_{\vec{k}}^{+}|^{2}e^{-i\vec k\cdot \vec r}\cr
995: &=&{1\over 2\pi^2}\int dk |{\vec{k}}|^2 |\delta T_{\vec{k}}^{+}|^{2} {\sin(|{\vec{k}}||{\vec{r}}|)\over
996: |{\vec{k}}||{\vec{r}}|}.
997: \label{eq:57}
998: \eeqa
999: In the following, we will always set $\vec r=0$ and consider
1000: \begin{equation}
1001: _{in}\langle 0|\delta{\sigma_{T}}^{I}(t,\vec{x})^{2}|0\rangle_{in}=\int{\frac{d^{3}k}{(2\pi)^{3}}|
1002: \delta{T_{\vec{k}}}^{+}|^{2}},
1003: \label{eq:57A}
1004: \end{equation}
1005: since this is sufficient for determining the fluctuation spectrum.
1006: The result for general $\vec r$ can be obtained
1007: by multiplying the integrand by $\sin(|\vec k||\vec r|)/|\vec k||\vec r|$.
1008: 
1009: Equation~\eq{eq:57A} can be computed for any time $t$ in Region I using the
1010: expression given in~\eq{eq:49}. However, for our purposes, it is most
1011: illuminating to evaluate it in the distant past. Inserting expression~\eq{eq:48},
1012: we find that the correlation function~\eq{eq:57A} becomes
1013: \begin{equation}
1014: _{in}\langle 0|{\delta\sigma_{T}}^{I}(t,\vec x)^{2}|0\rangle _{in}=f(t)\int\frac{d^3k}{(2\pi)^{3}}
1015: \frac{1}{2E_{\vec{k}}}
1016: \label{eq:58}
1017: \end{equation}
1018: where
1019: \begin{equation}
1020: f(t)=2e^{2\phi_{0}}e^{2Qt}.
1021: \label{eq:59}
1022: \end{equation}
1023: For ${\vec{k}}^{2}\gg Q^{2}$, the momentum regime~\eq{eq:40} in which we
1024: are working, expression~\eq{eq:58} becomes, to next to leading order,
1025: \begin{equation}
1026: _{in}\langle 0|{\delta\sigma_{T}}^{I}(t,\vec x)^{2}|0\rangle _{in}=f(t)\int\frac{d^3k}{(2\pi)^{3}}
1027: \frac{1}{|\vec{k}|}\left(\frac{1}{2}+\frac{Q^{2}/2}{2|\vec{k}|^{2}}\right).
1028: \label{eq:60}
1029: \end{equation}
1030: The momentum dependence of the first term of the integrand is simply that of zero-point fluctuations in Minkowski space. On the other hand,
1031: the second term corresponds precisely to a scale invariant fluctuation
1032: spectrum. Note, however, that since $\vec{k}^{2} \gg Q^{2}$, the second term is subdominant to the Minkowski fluctuations. The
1033: time-dependent factor $f(t)$ is not canonical. To
1034: understand its origin, we note that the kinetic energy term for $\sigma_{T}$
1035: in the Lagrangian~\eq{eq:2} is not canonically normalized in the
1036: gravitational background given in~\eq{eq:7} and~\eq{eq:8}. This kinetic
1037: energy term can be canonically normalized by defining a new scalar field
1038: ${\Sigma_{T}}^{I}$ as
1039: \begin{equation}
1040: {\Sigma_{T}}^{I}=e^{-\phi_{0}} |\sinh(2Qt)|^{1/2}\sigma_{T}.
1041: \label{eq:61}
1042: \end{equation}
1043: Note that in the far past this expression becomes
1044: \begin{equation}
1045: {\Sigma_{T}}^{I}=f(t)^{-1/2}\sigma_{T}, \qquad t\longrightarrow -\infty,
1046: \label{eq:62}
1047: \end{equation}
1048: where $f(t)$ is given in~\eq{eq:59}. It follows from this and~\eq{eq:60} that
1049: at early times
1050: \begin{equation}
1051: _{in}\langle 0|{\delta
1052: \Sigma_{T}}^{I}(t,\vec x)^{2}|0\rangle _{in}=\int\frac{d^3k}{(2\pi)^{3}}\frac{1}
1053: {|\vec{k}|}\left(\frac{1}{2}+\frac{Q^{2}/2}{|\vec{k}|^{2}}\right).
1054: \label{eq:63}
1055: \end{equation}
1056: We conclude that the factor $f(t)$
1057: in the correlation function~\eq{eq:60} is simply a scale factor that can be
1058: absorbed by a field redefinition or not, depending on taste. This concludes
1059: our analysis of quantum fluctuations in Region I.
1060: 
1061: 
1062: 
1063: We now want to discuss the quantum fluctuations of $\delta \sigma_{T}$ in
1064: Region II, that is, for conformal time in the positive interval $0<t<+\infty$.
1065: Note that the fluctuation equation~\eq{eq:37} is independent of which region
1066: is being considered. It follows that the analysis of quantum fluctuations in
1067: Region II is essentially identical to that in Region I. For this reason, we
1068: will simply present our results. To begin with, the fluctuations $\delta
1069: T_{\vec{k}}^{+}$ and $\delta T_{\vec{k}}^{-}$ given in~\eq{eq:49}
1070: and~\eq{eq:51} respectively remain a complete set of solutions of
1071: equation~\eq{eq:37}. It follows that any quantum fluctuation in Region II can
1072: be written as
1073: \begin{equation}
1074: {\delta \sigma_{T}}^{II}=\int\frac{d^3k}{(2\pi)^{3/2}}({a_{\vec{k}}}^{II}
1075: \delta T_{\vec{k}}^{+}(t)e^{i \vec{k}\cdot\vec{x}}+{a_{\vec{k}}}^{II
1076: \dagger} \delta T_{\vec{k}}^{-}(t)e^{-i \vec{k}\cdot\vec{x}}),
1077: \label{eq:64}
1078: \end{equation}
1079: where ${a_{\vec{k}}}^{II}$ and ${a_{\vec{k}}}^{II \dagger}$ satisfy the
1080: canonical commutation relations
1081: \begin{equation}
1082: [{a_{\vec{k}}}^{II},{a_{\vec{k'}}}^{II \dagger}]=\delta^{3}(\vec{k}-\vec{k'}),
1083: \qquad [{a_{\vec{k}}}^{II}, {a_{\vec{k'}}}^{II}]=[{a_{\vec{k}}}^{II \dagger},
1084: {a_{\vec{k'}}}^{II \dagger}]=0.
1085: \label{eq:65}
1086: \end{equation}
1087: The vacuum state is then defined as the normalized state $|0\rangle _{out}$
1088: satisfying
1089: \begin{equation}
1090: {a_{\vec{k}}}^{II}|0\rangle _{out}=0
1091: \label{eq:66}
1092: \end{equation}
1093: for all momenta $\vec{k}$. Again, there are many objects that one may wish to
1094: compute at this point. For example, in analogy with Region I, we find that in
1095: the distant future
1096: \begin{equation}
1097: _{out}\langle 0|{\delta
1098: \sigma_{T}}^{II}(t,\vec x)^{2}|0\rangle _{out}=g(t)\int\frac{d^3k}{(2\pi)^{3}}\frac{1}
1099: {|\vec{k}|}\left(\frac{1}{2}+\frac{Q^{2}/2}{|\vec{k}|^{2}}\right),
1100: \label{eq:67}
1101: \end{equation}
1102: where
1103: \begin{equation}
1104: g(t)=2e^{2\phi_{0}}e^{-2Qt}.
1105: \label{eq:68}
1106: \end{equation}
1107: As discussed previously, the factor of g(t) arises from the
1108: non-canonical normalization
1109: of the $\sigma_{T}$ kinetic energy term in~\eq{eq:2} with respect to the geometric
1110: background~\eq{eq:7} and~\eq{eq:8}. Proper normalization of this term can be
1111: achieved by defining
1112: \begin{equation}
1113: {\Sigma_{T}}^{II}=e^{-\phi_{0}}|\sinh(2Qt)|^{1/2}{\sigma_{T}}^{II}
1114: \label{eq:69}
1115: \end{equation}
1116: for any positive conformal time $t$. Note that in the far future this relation
1117: becomes
1118: \begin{equation}
1119: {\Sigma_{T}}^{II}=g(t)^{-1/2}{\sigma_{T}}^{II}, \qquad t\longrightarrow +\infty,
1120: \label{eq:70}
1121: \end{equation}
1122: where $g(t)$ is given in~\eq{eq:68}. In terms of this canonically normalized scalar
1123: field, the fluctuation spectrum~\eq{eq:67} becomes
1124: \begin{equation}
1125: _{out}\langle 0|{\delta
1126: \Sigma_{T}}^{II}(t,\vec x)^{2}|0\rangle _{out}=\int\frac{d^3k}{(2\pi)^{3}}\frac{1}{|\vec{k}|}
1127: \left(\frac{1}{2}+\frac{Q^{2}/2}{|\vec{k}|^{2}}\right).
1128: \label{eq:71}
1129: \end{equation}
1130: Therefore, the $g(t)$ factor in the
1131: two-point correlation function~\eq{eq:67} is simply a scale factor. It
1132: can be absorbed or not, depending on taste. This concludes our analysis of
1133: quantum fluctuations in Region II.
1134: 
1135: 
1136: As stated earlier, because of the curvature singularity at $t=0$, there is no
1137: classical relationship between Region I and Region II. Each represents an
1138: independent cosmology. The same is true for the quantum fluctuations that we
1139: have just discussed. The creation operators
1140: ${a_{\vec{k}}}^{I \dagger}$ acting on the vacuum $|0\rangle _{in}$ create a
1141: Hilbert space of states ${\cal{H}}^{I}$ representing the quantum theory in the
1142: Big Crunch geometry of Region I. Similarly, the operators ${a_{\vec{k}}}^{II \dagger}$
1143: acting on $|0\rangle _{out}$ create a Hilbert space
1144: ${\cal{H}}^{II}$ representing the quantum theory of the Big Bang geometry of
1145: Region II. A priori, there is absolutely no relation between ${\cal{H}}^{I}$
1146: and ${\cal{H}}^{II}$.
1147: 
1148: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1149: \subsection{Relating the Big Bang to the Big Crunch}\label{sub:BBBC}
1150: 
1151: It is the thesis of Ekpyrotic cosmology that the Universe did not begin at the
1152: Big Bang. Rather, it  had a prior history, connected to our present geometry
1153: via some catastrophic event. In the Big Crunch/Big Bang versions of Ekpyrotic
1154: theory, this catastrophic event is a spacetime singularity at $t=0$. This
1155: singularity is of the type found in the curvature scalar in Region I and Region II
1156: as $t\rightarrow -0$ and $t\rightarrow +0$ respectively. We will, therefore,
1157: within the context of the theory described by~\eq{eq:2}, construct a Big
1158: Crunch/Big Bang scenario by connecting Region I and Region II classically at
1159: the singular point $t=0$. Having done this, one must also specify a relation
1160: between the quantum theories on these two regions. The most naive approach,
1161: and the one we will adopt in this section, is to identify the two Hilbert
1162: spaces. That is, assume that
1163: \begin{equation}
1164: {\cal{H}}^{I}={\cal{H}}^{II} \equiv {\cal{H}}.
1165: \label{eq:72}
1166: \end{equation}
1167: One consequence of this is that the creation/annihilation operators
1168: ${a_{\vec{k}}}^{I}$,${a_{\vec{k}}}^{I
1169: \dagger}$ and ${a_{\vec{k}}}^{II}$,${a_{\vec{k}}}^{II \dagger}$ all act on the
1170: same Hilbert space ${\cal{H}}$. In ``normal'' quantum field theory, that is,
1171: when there is no geometric singularity or event horizon separating the past
1172: from the future, the ``out'' creation/annihilation operators are linearly
1173: related to the ``in'' creation/annihilation operators via a so-called
1174: Bogolubov transformation. We will assume that the same is true in our theory,
1175: despite the existence of a curvature singularity at $t=0$. That is, we
1176: postulate that
1177: \begin{equation}
1178: \left(\begin{array}{c} {a_{\vec{k}}}^{II \dagger} \\ {a_{-\vec{k}}}^{II} \end{array}
1179: \right)
1180: =\left(\begin{array}{cc} X^{*}(\vec k) & Y^{*}(\vec k) \\ Y(-\vec k) & X(-\vec k) \end{array}\right)
1181: \left(\begin{array}{c} {a_{\vec{k}}}^{I \dagger} \\ {a_{-\vec{k}}}^{I} \end{array}
1182: \right).
1183: \label{eq:73}
1184: \end{equation}
1185: The complex Bogolubov coefficients $X$ and $Y$ are not completely
1186: independent. They are constrained by the requirement that the
1187: Region I and Region II creation/annihilation operators continue to
1188: satisfy the canonical commutation relations~\eq{eq:55} and~\eq{eq:65}
1189: respectively. It follows that
1190: \begin{equation}
1191: |X|^{2}-|Y|^{2}=1.
1192: \label{eq:74}
1193: \end{equation}
1194: This is the only constraint on these coefficients. However, for simplicity, we
1195: will further assume that
1196: \be\label{assume}
1197: X(\vec k)=X(-\vec k),\ \ Y(\vec k)=Y(-\vec k).
1198: \ee
1199: Relaxing these assumptions will not change any of our conclusions.
1200: 
1201: Having postulated the relations~\eq{eq:72} and~\eq{eq:73}, one can now
1202: compute correlation functions that were not defined separately in Region I and
1203: Region II. Specifically, we want to calculate the two-point function
1204: \begin{equation}
1205: _{in}\langle 0|{\delta \sigma_{T}}^{II}(t,\vec x)^{2}|0\rangle _{in},
1206: \label{eq:75}
1207: \end{equation}
1208: where ${\delta \sigma_{T}}^{II}$ is the Region II field operator given in~\eq{eq:64},
1209: whereas $|0\rangle _{in}$ is the Region I vacuum defined in~\eq{eq:56}. This is
1210: easily accomplished using~\eq{eq:55},~\eq{eq:56},~\eq{eq:64},~\eq{eq:73}
1211: and~\eq{eq:74}. The result is
1212: $$
1213: _{in}\langle 0|{\delta
1214: \sigma_{T}}^{II}(t,\vec x)^{2}|0\rangle _{in}=
1215: $$
1216: \be\label{eq:76}
1217: =\int\frac{d^3k}{(2\pi)^{3}}
1218: \left((1+2|Y|^{2})|\delta T_{\vec{k}}^{+}|^{2}+
1219: XY\delta
1220: T_{\vec{k}}^{+2}+
1221: X^{*}Y^{*}\delta T_{\vec{k}}^{-2}\right),
1222: \ee
1223: with the fluctuations $\delta T_{\vec{k}}^{+}$ and $\delta
1224: T_{\vec{k}}^{-}$ given by~\eq{eq:49} and~\eq{eq:51} respectively.
1225: Of particular physical importance is the
1226: form of this correlation function in the distant future.
1227: As $t\rightarrow +\infty$, expression~\eq{eq:76} becomes
1228: \begin{equation}
1229: _{in}\langle 0|{\delta
1230: \sigma_{T}}^{II}(t,\vec x)^{2}|0\rangle _{in}=g(t)\int\frac{d^3k}{(2\pi)^{3}}\frac{1}
1231: {|\vec{k}|}\left(\frac{1}{2}+\frac{Q^{2}/2}{2|\vec{k}|^{2}}\right)\Delta(\vec{k},t),
1232: \label{eq:77}
1233: \end{equation}
1234: where the scale factor $g(t)$ is given in~\eq{eq:68} and
1235: \begin{equation}
1236: \Delta(\vec{k},t)=1+2|Y|^{2}+XYe^{-2iE_{\vec{k}}t}+X^{*}Y^{*}e^{2iE_{\vec{k}}t}.
1237: \label{eq:78}
1238: \end{equation}
1239: Written in terms of the scalar field ${\Sigma_{T}}^{II}$ defined
1240: in~\eq{eq:69}, and using~\eq{eq:70}, this simplifies to
1241: \begin{equation}
1242: _{in}\langle 0|{\delta\Sigma_{T}}^{II}(t,\vec x)^{2}|0\rangle _{in}=\int\frac{d^3k}{(2\pi)^{3}}
1243: \frac{1}{|\vec{k}|}\left(\frac{1}{2}+\frac{Q^{2}/2}{2|\vec{k}|^{2}}\right)\Delta(\vec{k},t).
1244: \label{eq:79}
1245: \end{equation}
1246: 
1247: 
1248: 
1249: 
1250: 
1251: We want to compare these results to another correlation function, namely
1252: \begin{equation}
1253: _{in}\langle 0|{\delta \sigma_{T}}^{I}(t,\vec x)^{2}|0\rangle _{in}.
1254: \label{eq:80}
1255: \end{equation}
1256: Note that, in this case, both the field operator ${\delta \sigma_{T}}^{I}$ and the
1257: vacuum $|0\rangle _{in}$ are the Region I quantities defined in~\eq{eq:53}
1258: and~\eq{eq:56} respectively. For arbitrary time $t$, this two-point function
1259: was evaluated in~\eq{eq:57} and its functional form in the limit
1260: $t\rightarrow -\infty$ presented in~\eq{eq:60}. Finally, in terms of the
1261: scalar field ${\Sigma_{T}}^{I}$ defined in~\eq{eq:61}, and
1262: using~\eq{eq:62}, the $t\rightarrow -\infty$ limit of this
1263: correlation function becomes
1264: \begin{equation}
1265: _{in}\langle 0|{\delta \Sigma_{T}}^{I}(t,\vec x)^{2}|0\rangle _{in}=\int\frac{d^3k}{(2\pi)^{3}}
1266: \frac{1}{|\vec{k}|}\left(\frac{1}{2}+\frac{Q^{2}/2}{2|\vec{k}|^{2}}\right).
1267: \label{eq:81}
1268: \end{equation}
1269: Although these expressions were first derived strictly within the context of
1270: Region I, they remain valid for the complete Hilbert space ${\cal{H}}$. First
1271: note that, for arbitrary conformal time $t$, correlation functions~\eq{eq:76}
1272: and~\eq{eq:57} are identical only if the complex Bogolubov coefficient $Y=0$.
1273: This remains true when comparing the $t\rightarrow +\infty$ limit of
1274: $_{in}\langle 0|{\delta \sigma_{T}}^{II}(t,\vec x)^{2}|0\rangle _{in}$ given in~\eq{eq:77} to the
1275: $t\rightarrow -\infty$ limit of $_{in}\langle 0|{\delta
1276: \sigma_{T}}^{I}(t,\vec x)^{2}|0\rangle _{in}$ presented in~\eq{eq:56}. The simplest
1277: comparison can be made between the future correlation function~\eq{eq:79} and
1278: the past correlation function~\eq{eq:81}, since the irrelevant scale factors
1279: have been removed. As stated earlier, the form of the argument of
1280: the momentum integral in~\eq{eq:81} is that of Minkowski fluctuations plus a subdominant scale invariant contribution in the far past.
1281: However, the
1282: $\Delta(\vec{k},t)$ factor in~\eq{eq:79} potentially modifies the
1283: fluctuation spectrum in the far future. If coefficient $Y=0$, then it follows
1284: from~\eq{eq:78} that $\Delta=1$ and one again obtains the spectrum \eq{eq:81}.
1285: However, if $Y\neq0$, then $\Delta\neq1$ and the fluctuation
1286: spectrum is modified. What are the physical consequences of this?
1287: 
1288: 
1289: 
1290: A central assertion of Ekpyrotic Big Crunch/Big Bang theories is that a
1291: scale invariant fluctuation spectrum is generated in the Big Crunch
1292: geometry prior to the singularity which is then transmitted, without
1293: modification, to the Big Bang geometry after the singularity. It is these scale
1294: invariant fluctuations that are assumed to account for the observed
1295: fluctuations in the cosmic microwave background. But is this true? Or is the
1296: fluctuation spectrum modified by the presence of the singularity? There has
1297: been considerable controversy regarding this, with some authors concluding
1298: that the spectrum is transmitted unchanged, some authors claiming it is
1299: greatly modified and further literature showing that this question is
1300: ambiguous as posed, requiring more physics input to uniquely resolve it. All
1301: of this literature has attempted to confront this issue by imposing explicit boundary
1302: conditions to match the incoming and outgoing wavefunctions at, or near,
1303: the singularity.
1304: However, attempting to study the vicinity of a singularity is difficult since
1305: one expects short distance effects to greatly modify the geometry and the
1306: physics. This is the source of the ambiguities. Our point of view is
1307: different. We see that the question of the persistence of the
1308: spectrum through the singularity is precisely expressed by whether or not the
1309: Bogolubov coefficient $Y$ vanishes. If it vanishes, the
1310: spectrum is transmitted through the singularity into the future. If it does
1311: not vanish, the spectrum is modified after the Big Bang. We
1312: conclude that to study this problem, one must, unambiguously, compute the
1313: Bogolubov coefficient $Y$. But how? One could, of course, attempt to compute
1314: $Y$ by matching the wavefunctions at the singularity. But, as we have said,
1315: this process would be ambiguous. A far more concrete way would be to compute
1316: $Y$ directly from string theory, where at least for certain types of singularities
1317: global wavefunctions can be unambiguously defined.
1318: Given such globally defined wavefunctions, one can perform calculations
1319: in the asymptotic past and future, far away from the singularity at $t=0$.
1320: This is the approach we will follow in the remainder of this paper.
1321: However, to compute the Bogolubov coefficient $Y$ in this manner, it is necessary
1322: to demonstrate that the field theory defined by~\eq{eq:2} is, in fact, the low
1323: energy four-dimensional effective theory of some specific string theory. This
1324: is indeed the case, as we will now show.
1325: 
1326: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1327: \subsection{Four-Dimensional Cosmology from String
1328: Theory}\label{sub:four}
1329: 
1330: In ($d+1$)-dimensional spacetime, string theory gives rise to a classical
1331: effective action of the form
1332: \begin{equation}
1333: S_{d+1}=\frac{1}{2\kappa_{d+1}^2}\int d^{d+1}x\sqrt{-g}e^{-2\Phi}(R_{d+1}+4g^{IJ}
1334: \partial_{I}\Phi\partial_{J}\Phi-2\Lambda),
1335: \label{eq:82}
1336: \end{equation}
1337: where $I,J=0,1,\dots,d$, $g_{IJ}$ and $\Phi$ are the ($d+1$)-dimensional string frame metric
1338: and dilaton respectively and all other massless and massive string modes have
1339: been set to zero. A positive tree level cosmological constant $\Lambda>0$
1340: will arise, for example, in supercritical type II string theories in
1341: \begin{equation}
1342: d+1=26,42,58,\dots
1343: \label{eq:83}
1344: \end{equation}
1345: dimensions \cite{Antoniadis:1988vi}.
1346: 
1347: Let us now assume the existence of solutions of the string
1348: equations of motion with metrics of the form
1349: \begin{equation}
1350: g_{IJ}dx^{I}dx^{J}=g'_{\mu\nu}dx^{\mu}dx^{\nu}+e^{2\sigma_{T}'}\delta_{ab}
1351: dx^{a}dx^{b}+e^{2\sigma_{R}'}(dx^{d})^{2},
1352: \label{eq:84}
1353: \end{equation}
1354: where $\mu,\nu=0,1,2,3$, indices $a,b=4,\dots,d-1$ and $g_{\mu\nu}'$, $\sigma_{T}'$
1355: and $\sigma_{R}'$ are all functions of the four-dimensional coordinates
1356: $x^{\mu}$, $\mu=0,1,2,3$ only. In addition, we will take the dilaton $\Phi$ to
1357: be a function only of these four-dimensional coordinates. The coordinates
1358: $x^{a}$
1359: parameterize a ($d-4$)-torus with a reference radius $r_{0}$. This radius will typically be chosen to be the string scale, $\sqrt{\alpha'}$, or perhaps a few orders of magnitude larger.
1360: Similarly, the $d$-direction is compactified
1361: on a circle or an interval, either having a reference radius which, for simplicity, we also choose to be $r_{0}$.  For length scales much larger than this radius,  all heavy modes
1362: will decouple and we will arrive at a four-dimensional effective theory
1363: describing the light modes. These light modes will necessarily include $g_{\mu\nu}'$, $\sigma_{T}'$, $\sigma_{R}'$ and $\Phi$. However, one should also make sure that there are no additional ``stringy'' light modes, such as winding modes around one of the internal circles.
1364: For the moment, let us ignore this important subtlety, returning to it at the end of this section.
1365: The associated action is most easily computed
1366: from~\eq{eq:82} by Weyl rescaling the four-dimensional metric as
1367: \begin{equation}
1368: g_{\mu\nu}'=\Omega^{2} g_{\mu\nu},
1369: \label{eq:85}
1370: \end{equation}
1371: where
1372: \begin{equation}
1373: \Omega^{2}=e^{2\Phi_{4}'}
1374: \label{eq:86}
1375: \end{equation}
1376: and
1377: \begin{equation}
1378: \Phi_{4}'=\Phi-\frac{1}{2}(\sigma_{R}'+(d-4)\sigma_{T}').
1379: \label{eq:87}
1380: \end{equation}
1381: Then, making the field redefinitions
1382: \begin{equation}
1383: \phi=2\Phi_{4}', \qquad \sigma_{T}=\sqrt{2(d-4)}\sigma_{T}', \qquad
1384: \sigma_{R}=\sqrt{2}\sigma_{R}',
1385: \label{eq:88}
1386: \end{equation}
1387: defining
1388: \begin{equation}
1389: \frac{1}{{\kappa_{4}}^{2}}=\frac{V_{T}V_{R}}{{\kappa_{d+1}}^{2}}
1390: \label{eq:89}
1391: \end{equation}
1392: where $V_{T}$ and $V_{R}$ are the volumes of the $(d-4)$-torus and the
1393: $d$-direction circle/interval respectively and dropping all higher derivative
1394: terms, we find that the low energy action is given by
1395: \begin{equation}
1396: S=\frac{1}{2{\kappa_{4}}^{2}}\int d^{4}x\sqrt{-g}{\cal{L}},
1397: \label{eq:90}
1398: \end{equation}
1399: with
1400: \begin{equation}
1401: {\cal{L}}=R-\frac{1}{2}g^{\mu\nu}\partial_{\mu}\phi\partial_{\nu}\phi
1402: -\frac{1}{2}g^{\mu\nu}\partial_{\mu}\sigma_{T}\partial_{\nu}\sigma_{T}
1403: -\frac{1}{2}g^{\mu\nu}\partial_{\mu}\sigma_{R}\partial_{\nu}\sigma_{R}
1404: -2e^{\phi}\Lambda.
1405: \label{eq:91}
1406: \end{equation}
1407: Note that this is exactly the four-dimensional theory introduced in~\eq{eq:1}
1408: and~\eq{eq:2} if we set\footnote{In the simple supercritical
1409: string models we are considering, the cosmological constant $\Lambda$ is of order the
1410: string scale, which is at odds with our assumption \eq{eq:40} that
1411: $Q$ is small compared to the momenta of interest. However, in the critical ``cigar geometry'' mentioned in the Introduction,
1412: it is possible to choose $Q$ to be arbitrarily small. For this reason, we will simply ignore this issue in our analysis.}
1413: \begin{equation}
1414: Q=\sqrt{\frac{\Lambda}{2}}.
1415: \label{eq:6}
1416: \end{equation}
1417: Recall, however, that this result is predicated on the assumption that there are no additional
1418: stringy light modes. We must now check to see under what conditions this will be the case.
1419: 
1420: 
1421: To analyze this, we must be more careful in our discussion of decoupling. It is clear
1422: from~\eq{eq:84} that, in the string frame \eq{eq:82}, the physical radii of the $(d-4)$-torus
1423: and the $d$-direction circle are
1424: \begin{equation}
1425: e^{\sigma_{T}'}r_{0}, \qquad e^{\sigma_{R}'}r_{0}
1426: \label{eq:A}
1427: \end{equation}
1428: respectively. From~\eq{eq:3} and~\eq{eq:88} we see that, for the background geometries of
1429: interest in this paper, $\sigma_{T}'$ vanishes and, hence, the physical $(d-4)$-torus
1430: radius is time-independent and fixed at $r_{0}$. On the other hand, it follows from~\eq{eq:4}
1431: and~\eq{eq:88} that the physical radius of the $d$-direction circle is time-dependent and given by
1432: \begin{equation}
1433: r_{R}(t)=|\tanh(Qt)|r_{0}.
1434: \label{eq:B}
1435: \end{equation}
1436: For very early and late times, $r_{R}(t)$ approaches $r_{0}$. Note, however, that
1437: \begin{equation}
1438: r_{R}(t)\longrightarrow 0, \qquad t\longrightarrow \pm0.
1439: \label{eq:C}
1440: \end{equation}
1441: Therefore, we must be concerned that stringy modes winding around the $d$-direction circle
1442: will become very light as $t\rightarrow \pm0$, thus substantially changing the low energy
1443: theory given in~\eq{eq:90} and~\eq{eq:91}. Can we estimate the time regime for which this
1444: effective theory is no longer valid? To do this, note that the winding modes around the
1445: $d$-direction circle typically have a mass
1446: \begin{equation}
1447: m_{winding}\cong \frac{r_{R}(t)}{\alpha'}.
1448: \label{eqD}
1449: \end{equation}
1450: It then follows from~\eq{eq:B}, and the fact that $r_{0}$ is of the order of the string
1451: scale or a few orders of magnitude larger, that the effective field theory~\eq{eq:90}
1452: and~\eq{eq:91} will be valid for momenta $\vec k$ satisfying
1453: \be\label{valid}
1454: \vec k^2\ll{r_0^2\tanh^2(Qt)\over(\alpha')^2}.
1455: \ee
1456: Since we are interested in momenta satisfying%
1457: \footnote{Massless modes with momenta below this scale
1458: correspond to growing modes and were discussed in section~5 of
1459: \cite{Craps:2002ii}. They lead to infrared divergences in string perturbation
1460: theory which will be ignored in this paper but would be interesting
1461: to understand better. We are not aware of any direct relation to the
1462: divergences in string perturbation theory mentioned at the beginning of
1463: section~\ref{sec:string}.}
1464: \be\label{momQ}
1465: Q^2< \vec k^2,
1466: \ee
1467: we have to make sure that there is a window of momenta satisfying both \eq{valid} and
1468: \eq{momQ}. Since we imagine $Q$ to be much below the string scale, $\alpha'Q^2\ll1$,
1469: and for $r_0^2\cong \alpha'$, such a window exists as long as
1470: \begin{equation}
1471: t^2\gg\alpha',
1472: \label{eq:E}
1473: \end{equation}
1474: that is, as long as we stay more than a few string times away from the singularity.
1475: Note that this  does not alter the discussion and
1476: conclusions associated with the theory given in~\eq{eq:1} and~\eq{eq:2}. To begin with,
1477: that theory is singular at the origin , so we could not really discuss the
1478: $t\rightarrow \pm0$ regime, nor did we. Indeed, the stringy effects that descend
1479: from~\eq{eq:82} in the time region~\eq{eq:E} could conceivably ``regulate'' this
1480: singularity, making the theory well-defined at the origin. We will not explore
1481: this possibility in this paper. Henceforth, we will simply recall that the low
1482: energy theory~\eq{eq:1},~\eq{eq:2} or, equivalently, ~\eq{eq:90},~\eq{eq:91} is
1483: not valid too close to the singularity.
1484: 
1485: 
1486: Subject to this caveat, we conclude that the four-dimensional cosmology described in this
1487: section descends from a class of string theories in $d+1$
1488: dimensions. We will now examine these string theories and their cosmology in detail,
1489: with the aim of eventually using them to compute the Bogolubov coefficients discussed above.
1490: 
1491: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1492: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1493: 
1494: \section{String Theory Cosmology}\label{sec:string}
1495: 
1496: As we have just shown, the four-dimensional action \eq{eq:1},~\eq{eq:2} is
1497: the low energy limit of the supercritical string theory \eq{eq:82}
1498: dimensionally reduced to four dimensions on a $(d-4)$-torus times a circle.
1499: The $(d+1)$-dimensional action~\eq{eq:82} is itself an effective theory, all
1500: higher $\alpha'$, string loop and non-perturbative effects being ignored. These
1501:  effects are important, and we will comment on them in the appropriate places
1502: later in this paper. However, we will begin by studying the classical  string
1503: action~\eq{eq:82} and, specifically, the $(d+1)$-dimensional cosmological
1504: solutions of its equations of motion. One such solution, the ``periodically
1505: identified generalized Milne solution'', or the ``generalized Milne orbifold''
1506: for short, was first presented in \cite{Craps:2002ii}.
1507: This turns out to be an exact solution of classical string theory, that is, it
1508: does not receive $\alpha'$ corrections \cite{TseytlinJackJonesPanvel}. We will see that, upon dimensional
1509: reduction to four dimensions, this solution includes the background spacetime
1510: studied in subsection~\ref{sub:background}
1511: This makes the generalized Milne orbifold an excellent context to try and answer the questions
1512: raised in the previous section.
1513: 
1514: 
1515: 
1516: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1517: \subsection{The Background Spacetime}\label{sub:string:background}
1518: 
1519: Our starting point is the following solution to the
1520: equations of motion associated with the classical string action \eq{eq:82}. The metric and dilaton are found to be
1521: \beqa\label{backgroundIIA}
1522: ds^2_{d+1}&=&{1\over Q^2}{du dv\over 1-uv}+
1523: \sum_{I=1}^{d-1}(dx^I)^2,\cr
1524: \Phi&=&-{1\over4}\log(1-uv)^2+\Phi_0.
1525: \eeqa
1526: Here
1527: \begin{equation}
1528: -\infty<u,v<\infty
1529: \label{eq:F}
1530: \end{equation}
1531: are global two-dimensional coordinates with the identification
1532: \be\label{uvid}
1533: (u,v)\sim (u e^{-2\pi Q r_{0}}, v e^{2\pi Qr_{0}}).
1534: \ee
1535: The parameter $Q$ is related to the cosmological constant $\Lambda$
1536: by expression \eq{eq:6}. The first term in the metric describes the
1537: generalized Milne orbifold \cite{Craps:2002ii}. The remaining terms
1538: describe  flat/torus-compactified $(d-1)$-dimensional space. The two-dimensional
1539: generalized Milne directions exhibit an intricate causal structure that is schematically represented in \fig{fig:CKR}.
1540: \begin{figure}
1541: \begin{center}
1542: \epsfig{file=cigar.eps, width=12.5cm}
1543: \end{center}
1544: \caption{The generalized Milne orbifold \cite{Craps:2002ii}.}\label{fig:CKR}
1545: \end{figure}
1546: 
1547: 
1548: In Regions I and II of \fig{fig:CKR},
1549: \begin{equation}
1550: uv<0
1551: \label{eq:G}
1552: \end{equation}
1553: and the identification \eq{uvid} is spacelike. These regions can also
1554: be described by the coordinates $t,x$ defined by
1555: \beqa\label{uvtx}
1556: u=\sinh (Qt) e^{-Qx}, \qquad v=-\sinh (Qt)\, e^{Qx}
1557: \eeqa
1558: where
1559: \be\label{xid}
1560: x\sim x+2\pi r_{0}.
1561: \ee
1562: In these coordinates, the metric and dilaton solutions become
1563: \beqa\label{backgroundIIAtx}
1564: ds^2_{d+1}&=&-dt^2+\tanh^2(Qt)dx^2+\sum_{I=1}^{d-1}(dx^I)^2,\cr
1565: \Phi&=&-\log\cosh(Qt)+\Phi_0.
1566: \eeqa
1567: For regions III and IV in \fig{fig:CKR},
1568: \begin{equation}
1569: 0<uv<1.
1570: \label{eq:H}
1571: \end{equation}
1572: The identification \eq{uvid} is now timelike. Thus, these regions contain closed timelike
1573: curves. In regions V and VI, one has
1574: \begin{equation}
1575: 1<uv
1576: \label{eq:I}
1577: \end{equation}
1578: and the identification again becomes spacelike.
1579: 
1580: 
1581: Region I describes a circle that starts out at some fixed radius
1582: $r_{0}$, collapses to zero size and then, in Region II, expands again to radius $r_{0}$.
1583: The Big Crunch/Big Bang singularity occurs when
1584: \begin{equation}
1585: uv=0.
1586: \label{eq:J}
1587: \end{equation}
1588: At this singularity, regions I and II touch the
1589: ``whisker'' \cite{Elitzur:2002rt} regions IV and III which, in turn,
1590: are connected to a second pair of asymptotic early and late time regions, VI and V respectively.
1591: In addition, the separation boundary between regions VI and IV and between regions V and III, defined by
1592: \begin{equation}
1593: uv=1,
1594: \label{eq:K}
1595: \end{equation}
1596: is singular. \fig{fig:CKR} is intended to give
1597: a rough idea of the structure of the asymptotic regions, but
1598: is less precise about the structure near the singularities.
1599: For example, the metric in Region VI is actually
1600: such that the radius of the circle is increasing
1601: as one moves in towards the singularity. The string
1602: coupling $g_s=\exp(\Phi)$ grows as well. In drawing  \fig{fig:CKR}, we have
1603: implicitly performed a T-duality locally on that region,
1604: to bring it to a form more like that of Region I.
1605: Close to the singularity at $uv=0$, where regions I, II, III and IV meet,
1606: the $u,v$ piece of \eq{backgroundIIA} reduces
1607: to the Milne orbifold, a two-dimensional Minkowski space with the points
1608: related by \eq{uvid} identified. See, for example, \cite{Cornalba:2002fi,
1609: Nekrasov:2002kf, Berkooz:2002je,Berkooz:2003bs}
1610: for recent discussions.
1611: Deep into regions I and II, that is, as $uv\rightarrow -\infty$, the dilaton becomes
1612: linear in time with the string coupling approaching zero.
1613: At $uv=1$, the separation
1614: between regions VI and IV and between regions V and III,
1615: both the curvature and
1616: the string coupling diverge.
1617: 
1618: 
1619: A priori, one would
1620: expect stringy, higher derivative corrections to \eq{eq:82}
1621: to become important near the singularities at $uv=0$ and $uv=1$.
1622: However, it was shown in  \cite{Craps:2002ii} that the background \eq{backgroundIIA}
1623: defines an exact superconformal field theory with the correct central charge, $\hat c=10$ \footnote{The central charge deficit
1624: of the $(u,v)$ directions is compensated by considering supercritical
1625: string theory with a positive cosmological constant.}. That is, background~\eq{backgroundIIA} is an exact solution of the string equations of motion to all orders in $\alpha'$.
1626: 
1627: 
1628: The coordinates $x^1,x^2,x^3$ range from $-\infty$ to $+\infty$.
1629: The coordinates $x^a, i=4,\ldots,d-1$ are taken to be periodic, $x^a\sim
1630: x^a+2\pi r_{0}$, so they describe a ($d-4$)-torus with constant radii $r_{0}$. These radii are chosen to be of the order of the string length or a few orders of magnitude larger.
1631: As mentioned previously, the coordinate $x$ defined in~\eq{uvtx} is also compact although, in string frame, the
1632: corresponding circle has a time-dependent radius $|\tanh(Qt)|r_{0}$.
1633: The coordinate $t$ will be the time coordinate of an observer living in regions I and II.
1634: 
1635: 
1636: We close this subsection by showing that in regions I and II of the $(d+1)$-dimensional cosmological solution given in~\eq{backgroundIIA}, there exists an effective
1637: four-dimensional description of this background . This effective solution is valid everywhere except very close to the Big Crunch/Big Bang singularity, where
1638: it is expected to break down due to additional
1639: light stringy modes. It turns out that  the dangerous modes are
1640: winding modes around the $x$ direction, which, as we have previously discussed,
1641: indeed become light near $t=0$. With this caveat in mind, we proceed
1642: to dimensionally reduce the metric and dilaton solutions in regions I and II
1643: from $(d+1)$-dimensions to four dimensions by compactifying them on the $(d-4)$-torus times a circle. Recall that in regions I and II the metric and dilaton can be written as in~\eq{backgroundIIAtx}. Using the definitions in subsection~\ref{sub:four}, we find that, at low energy, \eq{backgroundIIAtx} corresponds to
1644: \beqa\label{fourdeff}
1645: \sigma_T&=&0,\cr
1646: \sigma_R&=&\sqrt2\log|\tanh(Qt)|,\cr
1647: \phi&=&2\Phi_0+\log 2-\log|\sinh(2Qt)|,\cr
1648: g_{\mu\nu}&=&e^{-\phi}\eta_{\mu\nu},
1649: \eeqa
1650: which is exactly the four-dimensional background given in~\eq{eq:3},~\eq{eq:4},~\eq{eq:5} and~\eq{eq:7} of
1651: subsection~\ref{sub:background} if we identify
1652: \be\label{phizero}
1653: \phi_0=\Phi_0+{1\over2}\log2.
1654: \ee
1655: 
1656: It is crucial to note that \eq{backgroundIIAtx} only
1657: describes regions I and II of the $(d+1)$-dimensional
1658: spacetime \eq{backgroundIIA}. The full $(d+1)$-dimensional
1659: string background contains two additional asymptotic regions
1660: V and VI, as well as the intermediate regions III and IV.
1661: The four-dimensional spacetime \eq{fourdeff} is a low energy description of
1662: regions I and II only. However, as we have pointed out before, this
1663: four-dimensional effective description breaks down near the Big
1664: Crunch/Big Bang singularity, which is exactly where regions I
1665: and II are connected to the other regions. To understand what
1666: happens near the Big Crunch/Big Bang singularity, as well as in regions III, IV, V and VI, it is clearly necessary to use the full $(d+1)$-dimensional string theory background
1667: \eq{backgroundIIA}. This is precisely what we will do throughout the remainder of this paper.
1668: 
1669: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1670: \subsection{Fluctuations}\label{sub:string:fluctuations}
1671: 
1672: In this subsection, we study fluctuations around the background
1673: \eq{backgroundIIA}. The key ingredient in our discussion is that
1674: string theory allows one to determine globally defined
1675: wavefunctions, despite the singularities that prevent doing so in
1676: general relativity \cite{Craps:2002ii,Elitzur:2002rt}.
1677: The underlying reason is that the spacetime \eq{backgroundIIA}
1678: corresponds to an orbifold of a coset conformal field theory.%
1679: \footnote{For early work on applications of coset conformal field theories
1680: to cosmology, see for example \cite{coset,Kounnas:1992wc}.}
1681: There is a well-defined procedure to determine at least the
1682: ``untwisted'' globally defined vertex operators in such theories.
1683: The zero mode parts of these vertex operators are the globally
1684: defined wavefunctions we are interested in.
1685: 
1686: This procedure consists of two steps. In the first step, one
1687: determines the vertex operators of the coset conformal field theory,
1688: which in our case is $PSL(2,\Rbar)/U(1)$ at negative level.%
1689: \footnote{$PSL(2,\Rbar)/U(1)$ at positive level corresponds to a
1690: two-dimensional black hole geometry \cite{Witten:1991yr,Dijkgraaf:1991ba}.
1691: $PSL(2,\Rbar)/U(1)$ at negative level can be
1692: obtained from the black hole geometry by double Wick rotation \cite{Kounnas:1992wc}.
1693: It is described by the $u,v$ piece of \eq{backgroundIIA}, without the
1694: identification \eq{uvid}. In this spacetime, $u=v=0$ is a smooth
1695: point (the Big Crunch/Big Bang singularity only arises after
1696: the discrete identification \eq{uvid}), while there are
1697: singularities at $uv=1$.}
1698: One describes the coset conformal field theory as a gauged
1699: Wess-Zumino-Witten (WZW) model. The vertex operators of the
1700: ungauged $PSL(2,\Rbar)$ WZW model correspond to wavefunctions on
1701: the smooth $PSL(2,\Rbar)$ group manifold. These can be found in
1702: \cite{Vilenkin}.
1703: The vertex operators of the $SL(2,\Rbar)/U(1)$ WZW model can then
1704: be viewed as those vertex operators of the ungauged model that are
1705: invariant under the $U(1)$ gauge group. See for example
1706: \cite{Craps:2002ii, Elitzur:2002rt, Dijkgraaf:1991ba}.
1707: 
1708: The second step is the standard string theory orbifold procedure
1709: \cite{orb}.
1710: Roughly speaking, an orbifold is obtained from a
1711: covering space, in our case the $PSL(2,\Rbar)/U(1)$ coset spacetime at negative
1712: level,
1713: by identifying points related by the action
1714: of a discrete group. Here, the group is $\Zbar$ with the action
1715: \eq{uvid}, which turns a line into a circle (the circles visible in
1716: Fig.~\ref{fig:CKR}) and causes the Big Crunch/Big Bang singularity at
1717: $uv=0$. The untwisted vertex operators%
1718: \footnote{There are also twisted vertex operators, corresponding
1719: to strings winding around the circles, see \cite{Craps:2002ii}.}
1720: are those
1721: vertex operators on the covering space that are invariant under
1722: \eq{uvid}. From \eq{xid}, we see that this amounts to momentum
1723: quantization in the $x$
1724: direction. In this paper, we will only be
1725: interested in zero momentum in the $x$ direction since the $x$
1726: circle is part of the internal space. For the same reason, we will
1727: not be interested in winding modes except to note that they cause
1728: the effective four-dimensional description to break down near the
1729: Big Crunch/Big Bang.
1730: 
1731: We would like to comment on the relationship of our theory to another model,
1732: which has the same Big Crunch/Big Bang singularity. This is the Milne
1733: orbifold $(\Rbar^{(1,1)}/\Zbar) \times \Rbar^8$, where $\Zbar$ is generated by
1734: the boost transformation \eq{uvid} on two-dimensional Minkowski
1735: space and $\Rbar^8$ denotes eight additional flat directions, some
1736: of which may be compactified. The associated metric is
1737: \be\label{Milne}
1738: ds^2={1\over Q^2}dudv+\sum_{I=1}^8(dx^I)^2.
1739: \ee
1740: This spacetime is very similar to \eq{backgroundIIA} near the Big
1741: Crunch/Big Bang singularity $u=v=0$, but differs significantly
1742: from it away from this point, in particular in the structure of the
1743: additional regions. The Milne orbifold consists of four cones
1744: touching at $u=v=0$, that is, regions I, II, III and IV, but the
1745: radius of the circle of the cones grows to infinite size as one
1746: goes infinitely far away from the singularity. There are no
1747: regions V and VI. They can be thought of as having been pushed
1748: to infinity by focussing in on the manifold around $u=v=0$.
1749: 
1750: Wavefunctions on the Milne orbifold have been discussed from a
1751: string theory point of view in \cite{Nekrasov:2002kf} and have
1752: been used to compute string scattering amplitudes in
1753: \cite{Berkooz:2002je}. There is no coset CFT involved in this
1754: spacetime and the untwisted wavefunctions can be easily obtained from the
1755: string orbifold procedure. That is, consider wavefunctions on Minkowski space
1756: and demand that they be invariant under the action
1757: \eq{uvid} of the orbifold group. Actually, such invariant
1758: Minkowski space wavefunctions either grow or decay
1759: in regions III and IV. Because these regions are non-compact
1760: in the Milne orbifold, one, often implicitly, further restricts to
1761: wavefunctions that decay in those regions. There
1762: will be no such additional restriction in the spacetime
1763: \eq{backgroundIIA}, since regions III and IV do not extend
1764: to infinity there. As a consequence, more wavefunctions
1765: are necessary in our theory than one finds in the Milne orbifold.
1766: 
1767: Global wavefunctions for the Milne orbifold with regions III and
1768: IV omitted were presented in \cite{Tolley:2002cv} based on a
1769: construction that does not refer to string theory, and applied in
1770: \cite{Tolley:2003nx} to cosmology. The global wavefunctions agree
1771: with the restriction to regions I and II of the stringy
1772: wavefunctions discussed in the previous paragraph.
1773: Also, the vacuum state implicitly defined by the wavefunctions of
1774: \cite{Nekrasov:2002kf} and more explicitly used in \cite{Berkooz:2002je}
1775: corresponds to the vacuum state defined
1776: in \cite{Tolley:2002cv} upon deleting regions III and IV of
1777: the string theory spacetime (or from the point of view of \cite{Tolley:2002cv},
1778: upon adding those regions). It is of interest to see what happens to
1779: the wavefunctions of the generalized Milne orbifold upon taking the limit
1780: to the Milne orbifold. We analyze this in detail in subsection 2.4 and
1781: Appendix~C.
1782: 
1783: The fact that the wavefunctions that descend from a covering conformal field
1784: theory solution of string theory are globally defined on the associated
1785: orbifold is of fundamental importance for the results of this paper. For that
1786: reason, we outline in Appendix A, in more detail than discussed here, the procedure for
1787: constructing these global wavefuntions.
1788: In the remainder of this subsection, we will review and extend the
1789: results of \cite{Craps:2002ii} on global wavefunctions and quantum
1790: vacuum states in the spacetime \eq{backgroundIIA}.
1791: The global
1792: structure of the spacetime, in particular the presence of the
1793: additional asymptotic regions, will be shown to have important
1794: implications for the physics of the Big Crunch/Big Bang
1795: transition.
1796: 
1797: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1798: \subsubsection{Global Wavefunctions on the Generalized Milne Orbifold}
1799: \label{subsub:global}
1800: We restrict our discussion to those fluctuations that are relevant for the
1801: four-dimensional cosmology introduced in section~\ref{sec:fourd}. That is, the only
1802: non-zero momentum components are in the three non-compact
1803: space dimensions, and we ignore all winding and excited string
1804: modes. In particular, momentum and winding in the $x$ direction
1805: are set to zero,
1806: \be\label{pw}
1807: p=w=0
1808: \ee
1809: in the notation of \cite{Craps:2002ii}.
1810: The justification for this is that all the modes we ignore have masses
1811: on the order of the string scale, except very close to the Big Bang/Big Crunch singularity
1812: where winding modes around the $x$ direction become light, as we
1813: have previously mentioned.
1814: %By staying sufficiently far from this singularity and considering energies
1815: %and momenta much smaller than the string scale,
1816: %\be\label{momstringscale}
1817: %\vec k^2\ll{1\over \alpha'},
1818: %\ee
1819: %all heavy momentum and winding modes can safely be ignored.
1820: %Recall, however, that we must restrict our discussion to momenta satisfying
1821: %\be\label{momQ}
1822: %\vec k^2\gg Q^2,
1823: %\ee
1824: %where $Q$ is defined in~\eq{eq:6}, in order to be consistent with the
1825: %condition \eq{eq:40} imposed in section~\ref{sec:fourd}.
1826: 
1827: In subsection~\ref{sub:fluctuations}, we solved  the wave equation
1828: \eq{eq:37} for the fluctuations $\delta T$ defined in~\eq{eq:36}. Recall that from the point of view of the four-dimensional quantum field theory~\eq{eq:1} and~\eq{eq:2}, there were two classically independent regions, Region I and Region II.
1829: In each of these regions, we found the same two independent solutions of~\eq{eq:37}
1830: for a given momentum
1831: $\vec k$. These solutions are ${\delta T_{\vec{k}}}^{+}$ in~\eq{eq:49} and
1832: ${\delta T_{\vec{k}}}^{-}(= {\delta T_{\vec{k}}}^{+*}) $.
1833: Far from $t=0$, one solution, ${\delta T_{\vec{k}}}^{+}$, reduces to a positive
1834: frequency wave, while its conjugate has negative frequency. However, as we have
1835: shown in
1836: subsection~\ref{sub:string:background}, this four-dimensional theory arises as
1837: the low energy limit of the generalized Milne orbifold solution of the
1838: $(d+1)$-dimensional classical string action~\eq{eq:82}. The full string theory
1839: background \eq{backgroundIIA} has, in addition to regions I and II, the regions III,
1840: IV,  V and VI discussed in detail above. Therefore, in string theory, one must solve the
1841: $\delta T$ fluctuation equation in each of these six regions. Recall that
1842: the generalized Milne orbifold arises from a covering $PSL(2,\Rbar)$ manifold by
1843: constructing  the coset $PSL(2,\Rbar)/U(1)$ and then identifying points related by a
1844:  $\Zbar$ group action. It follows from this structure that one can find solutions of
1845: the $\delta T$ fluctuation equation on the generalized Milne orbifold by first finding
1846: solutions of the fluctuation equation on $PSL(2,\Rbar)$ and then restricting to those
1847: solutions that are invariant under the action of both $U(1)$ and $\Zbar$. Such
1848: solutions, since they are globally defined on $PSL(2,\Rbar)$, remain globally
1849: defined on the generalized Milne orbifold. That is, each such solution is defined
1850: in each of the six regions I, II, III, IV, V and VI. As explained in
1851: \cite{Craps:2002ii,Elitzur:2002rt}, there are
1852: four independent wavefunctions of this type for each momentum $\vec k$,
1853:  which we denote by
1854: \begin{equation}
1855: \label{Kuv}
1856: \K_{++,\vec{k}}(uv),\ \K_{+-,\vec{k}}(uv),\ \K_{-+,\vec{k}}(uv)\ {\rm and}\
1857: \K_{--,\vec{k}}(uv).
1858: \ee
1859: Note that, in addition to their $\vec{k}$ dependence, the argument of these
1860: functions is the coordinate $uv$. More precisely, they depend not only on $uv$
1861: but, also, on the region I, II, III, IV, V or VI. This latter dependence is suppressed
1862: in \eq{Kuv}. In regions I and II,
1863: it follows from~\eq{uvtx} that
1864: \begin{equation}
1865: uv=-\sinh^{2}(Qt).
1866: \label{eq:n1}
1867: \end{equation}
1868: Similarly, in regions V and VI we can write
1869: \begin{equation}
1870: 1-uv=-\sinh^{2}(Qt).
1871: \label{eq:n1bis}
1872: \end{equation}
1873: Hence, in these regions the wavefunctions are dependent on $t$,
1874: as we would expect them to be.
1875: The four independent wavefunctions in~\eq{Kuv} can be written as
1876: \begin{equation}
1877: \K_{\pm\pm,\vec{k}}=\N K_{\pm\pm,\vec{k}},
1878: \label{eq:n2}
1879: \end{equation}
1880: where $\N$ is a $\vec{k}$-dependent normalization constant given in~\eq{calN} and
1881: the expressions for each of the four functions $K_{\pm\pm, \vec{k}}$ in the
1882: six regions are given in Appendix B. That is, the independent wavefunctions
1883: in~\eq{Kuv} are explicitly known.
1884: %cite{Vilenkin, Elitzur:2002rt, Craps:2002ii}. Restricted
1885: %to region~$I$, for instance, $\K_{++}$ coincides with $\delta
1886: %T_{\vec k}^+$ (see \eq{eq:49}), $\K_{+-}$ vanishes, $\K_{-+}$ is a
1887: %linear combination of $\delta T_{\vec k}^+$ and $\delta T_{\vec k}^-$
1888: %(see \eq{eq:51}), and $\K_{--}$ coincides with $\delta T_{\vec k}^-$
1889: %(up to a phase).
1890: The $\K_{\pm\pm, \vec{k}}$ are defined such
1891: that they are orthonormal, with norm squared $\pm 1$, with respect
1892: to the appropriate generalization of the Klein-Gordon inner product
1893: \eq{eq:44}. The
1894: generalization is that $\Sigma$ should be a global ``Cauchy''
1895: surface, intersecting regions I and VI or II and V, which
1896: turn out to be equivalent. The $K_{\pm\pm,\vec{k}}$ given in Appendix A have the same norm
1897: up to a sign, so, by dividing out a common factor, we can obtain
1898: the normalized $\K_{\pm\pm,\vec{k}}$. For example, $\K_{++,\vec{k}}$ coincides
1899: with $\delta T_{\vec k}^+$ in Region I and vanishes in Region
1900: VI, so it indeed has unit Klein-Gordon norm.
1901: 
1902: 
1903: A key point of this paper is that the wavefunctions $\K_{\pm\pm,\vec{k}}$
1904: are globally
1905: defined on the generalized Milne orbifold. This follows from the fact that
1906: they descend from globally defined functions on the $PSL(2,\Rbar)$ covering
1907: space. The expressions for each of the four independent wavefunctions, in each
1908: of the six regions, are presented in Appendix B. It is, however, useful at
1909: this point to discuss several of these wavefunctions in more detail. First
1910: consider $\K_{++, \vec{k}}$. In Region I, it is found to be
1911: \begin{equation}
1912: \K_{++, \vec{k}}=\delta {T_{\vec{k}}}^{+},
1913: \label{eq:c1}
1914: \end{equation}
1915: where $ \delta {T_{\vec{k}}}^{+}$ is given in~\eq{eq:49}.
1916: This function is purely positive
1917: frequency in the asymptotic region $t\ll -1/Q$, while it diverges
1918: logarithmically near the Big Crunch singularity, $t\to -0$. In Region II,
1919: one has
1920: \begin{equation}
1921: \K_{++, \vec{k}}\propto (-z)^{-j-1}F(-j,-j;-2j;\frac{1}{z}),
1922: \label{eq:c2}
1923: \end{equation}
1924: where $j$ and $z$ are defined in~\eq{eq:50}.
1925: In this region, $\K_{++, \vec{k}}$ is
1926: purely positive frequency asymptotically and diverges logarithmically
1927: near the Big Bang as $t\to +0$. In the intermediate
1928: Region III,
1929: \begin{equation}
1930: \K_{++, \vec{k}}\propto (uv)^{j}F(-j,-j;1;1-\frac{1}{uv}),
1931: \label{eq:c3}
1932: \end{equation}
1933: which diverges logarithmically near the Big Crunch/Big Bang singularity at
1934: $uv=0$, while it approaches a constant near the $uv=1$ singularity that separates
1935: regions III and V. Similarly, in Region IV
1936: \begin{equation}
1937: \K_{++, \vec{k}}\propto (uv)^{-j-1}F(j+1,j+1;1;1-\frac{1}{uv}).
1938: \label{eq:c4}
1939: \end{equation}
1940: This diverges logarithmically
1941: near $uv=0$ and approaches a constant near $uv=1$. In
1942: Region V, $\K_{++, \vec{k}}$ is
1943: a mixture of positive and negative frequency components with equal amplitude in the
1944: asymptotic region $uv\gg 1/Q^2$, while it approaches a constant near $uv=1$.
1945: The exact expression in Region V is
1946: \begin{equation}
1947: \K_{++, \vec{k}}\propto F(-j,j+1;1;1-uv).
1948: \label{eq:c5}
1949: \end{equation}
1950: Finally, and importantly, in Region VI
1951: \begin{equation}
1952: \K_{++, \vec{k}}=0.
1953: \label{eq:c6}
1954: \end{equation}
1955: 
1956: The global behavior of $\K_{+-, \vec{k}}$ can be obtained by replacing
1957: $I\leftrightarrow VI$,
1958: $II\leftrightarrow V$ and $uv\leftrightarrow 1-uv$ in the above expressions for
1959: $\K_{++, \vec{k}}$. Of particular importance for us is the form of $\K_{+-,
1960: \vec{k}}$ in regions I and VI. We find that
1961: \begin{equation}
1962: \K_{+-, \vec{k}}=0
1963: \label{eq:c7}
1964: \end{equation}
1965: and
1966: \begin{equation}
1967: \K_{+-, \vec{k}}=\delta {T_{\vec{k}}}^{+},
1968: \label{eq:c8}
1969: \end{equation}
1970: in regions I and VI respectively. Similarly, the
1971: global structure of $\K_{--, \vec{k}}$ is obtained from that of $\K_{++,
1972: \vec{k}}$ by replacing $I
1973: \leftrightarrow II$, $V\leftrightarrow VI$ and ``positive frequency''
1974: $\leftrightarrow$ ``negative frequency''. Specifically, we will use the fact
1975: that
1976: \begin{equation}
1977: \K_{--, \vec{k}}=\delta {T_{\vec{k}}}^{+*}
1978: \label{eq:c9}
1979: \end{equation}
1980: and
1981: \begin{equation}
1982: \K_{--, \vec{k}}=0
1983: \label{eq:c10}
1984: \end{equation}
1985: in regions II and V respectively.
1986: Finally, the behavior of $\K_{-+, \vec{k}}$ is
1987: obtained from that of $\K_{--, \vec{k}}$ by replacing $I\leftrightarrow VI$,
1988: $II\leftrightarrow V$ and $uv\leftrightarrow 1-uv$. In particular,
1989: we find that
1990: \begin{equation}
1991: \K_{-+, \vec{k}}=0
1992: \label{eq:c11}
1993: \end{equation}
1994: in Region II, whereas
1995: \begin{equation}
1996: \K_{-+, \vec{k}}=\delta {T_{\vec{k}}}^{+*}
1997: \label{eq:c12}
1998: \end{equation}
1999: in Region V.
2000: 
2001: 
2002: 
2003: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2004: 
2005: \subsubsection{Quantization}\label{subsub:quantization}
2006: 
2007: We would now like to quantize the scalar fluctuations
2008: $\delta \sigma_{T}$ on the
2009: generalized Milne orbifold. Generically, this can be done by
2010: expanding
2011: \beqa\label{expand}
2012: &&\delta\sigma_T=\int{d^3k\over(2\pi)^{3/2}}\left(a_{1,\vec k}\K_{1,\vec k}(uv)
2013: e^{i\vec k
2014: \cdot\vec x}+a_{2,\vec k}\K_{2,\vec k}(uv)e^{i\vec k\cdot\vec x}\right.\cr
2015: &&\ \ \ \ \ \ \ \ \ \ \ \ \ \ \  \ \ \ \ \ \  \left.
2016: +a_{1,\vec k}^{\dagger}\K_{1,\vec k}^*(uv)e^{-i\vec k\cdot\vec x}
2017: +a_{2,\vec k}^{\dagger}\K_{2,\vec k}^*(uv)e^{-i\vec k\cdot\vec x}\right),
2018: \eeqa
2019: where $\{\K_{1,\vec k}, \K_{2,\vec k}\}$ is some appropriate pair of
2020: orthonormal wavefunctions.
2021: As compared to \eq{eq:53}, we have twice as
2022: many functions in the expansion \eq{expand}. The reason is that there are twice
2023: as many independent global wavefunctions on the generalized Milne orbifold
2024: as there are independent wavefunctions in the effective four-dimensional
2025: theory.
2026: We now impose the canonical commutation relations
2027: \be\label{cancomm}
2028:  [a_{i,\vec k},a_{j,\vec k'}^{\dagger}]=
2029: \delta_{ij}\delta^3(\vec k-\vec k'), \qquad [a_{i,\vec k},a_{j,\vec k'}]=
2030: [a_{i,\vec k}^{\dagger},a_{j,\vec k'}^{\dagger}]=0.
2031: \ee
2032: and define the vacuum state $|0\rangle$ by
2033: \be\label{vacuumstate}
2034: a_1|0\rangle=a_2|0\rangle=0.
2035: \ee
2036: 
2037: 
2038: To proceed, we now must ask the question: what is a natural vacuum state to choose, or equivalently,
2039: what is a natural set of wavefunctions $\{\K_{1,\vec k}, \K_{2,\vec k}\}$? In
2040: \cite{Craps:2002ii}, two natural vacua were defined and shown to be
2041: inequivalent. The first vacuum state corresponds to the
2042: choice
2043: $\K_{1,\vec{k}}=\K_{++, \vec{k}},\ K_{2, \vec{k}}=\K_{+-, \vec{k}}$. The
2044: associated fluctuation expansion is
2045: \beqa\label{expandin}
2046: &&\delta\sigma_T=\int{d^3k\over(2\pi)^{3/2}}\left(a^{I}_{\vec{k}}\K_{++,\vec
2047: k}(uv)
2048: e^{i\vec k
2049: \cdot\vec x}+a^{VI}_{\vec k}\K_{+-,\vec k}(uv)e^{i\vec k\cdot\vec x}\right.\cr
2050: &&\ \ \ \ \ \ \ \ \ \ \ \ \ \ \  \ \ \ \ \ \  \left.
2051: +a^{I \dagger}_{\vec{k}}\K_{++,\vec k}^*(uv)e^{-i\vec k\cdot\vec x}
2052: +a^{VI \dagger}_{\vec{k}}\K_{+-,\vec k}^{*}(uv)e^{-i\vec k\cdot\vec x}\right).
2053: \eeqa
2054: Recall that at early times, $\K_{++, \vec{k}}$ is purely positive frequency in
2055: Region I and, from~\eq{eq:c6}, vanishes
2056: in Region VI, and vice versa for the wavefunction
2057: $\K_{+-, \vec{k}}$. It follows that
2058: the vacuum state constructed from $a^{I}_{\vec{k}}$ and $a^{VI}_{\vec{k}}$
2059: would indeed be called empty
2060: by early time observers in regions I and VI. Therefore, we denote this state by
2061: $|0\rangle_{in}$.
2062: The second vacuum state, specified by $|0\rangle_{out}$, corresponds to the
2063: choice
2064: $\K_{1, \vec{k}}=\K^*_{--, \vec{k}}$ and $\K_{2, \vec{k}}=\K^*_{-+, \vec{k}}$
2065: and the associated expansion
2066: \beqa\label{expandout}
2067: &&\delta\sigma_T=\int{d^3k\over(2\pi)^{3/2}}\left(a^{II}_{\vec k}
2068: \K^*_{--,\vec k}(uv)
2069: e^{i\vec k
2070: \cdot\vec x}+a^{V}_{\vec k}\K^*_{-+,\vec k}(uv)e^{i\vec k\cdot\vec x}\right.\cr
2071: &&\ \ \ \ \ \ \ \ \ \ \ \ \ \ \  \ \ \ \ \ \  \left.
2072: +a^{II\dagger}_{\vec k}\K_{--,\vec k}(uv)e^{-i\vec k\cdot\vec x}
2073: +a^{V\dagger}_{\vec k}\K_{+-,\vec k}(uv)e^{-i\vec k\cdot\vec x}\right).
2074: \eeqa
2075: Since, at late times, $\K^*_{--, \vec{k}}$ is purely positive frequency in Region
2076: II and, from~\eq{eq:c10}, vanishes
2077: in Region V, and vice versa for $\K_{-+, \vec{k}}$, this state would similarly be called empty
2078: by late time observers in regions II and V.
2079: 
2080: The relation between $|0\rangle_{in}$ and $|0\rangle_{out}$ can be determined
2081: from
2082: \be\label{bogol}
2083: \pmatrix{\K_{--,\vec{k}}\cr \K_{-+,\vec{k}}\cr \K^*_{--, \vec{k}}\cr
2084: \K^*_{-+,\vec{k}}}=
2085: \pmatrix{A&C&0&B\cr
2086:          C&A&B&0\cr
2087:          0&B^*&A^*&C^*\cr
2088:          B^*&0&C^*&A^*}
2089: \pmatrix{\K^*_{++, \vec{k}}\cr \K^*_{+-, \vec{k}}\cr \K_{++, \vec{k}}\cr
2090: \K_{+-, \vec{k}}},
2091: \ee
2092: where the Klein-Gordon orthonormality of the wavefunctions implies that
2093: $A$, $B$ and $C$ satisfy the conditions
2094: \begin{equation}
2095: |A|^2+|C|^2-|B|^2=1, \qquad AC^*+A^*C=0.
2096: \label{consistency}
2097: \end{equation}
2098: For the special case \eq{pw} we are considering, one finds
2099: \beqa\label{ABC}
2100: A&=&-1,\cr
2101: B=C&=&{1\over i\sinh\left({\pi E_{\vec{k}}\over 2Q}\right)}.
2102: \eeqa
2103: Note that $B$ is a function of $\vec{k}$ and that $B(\vec{k})=B(-\vec{k})$.
2104: Then,~\eq{bogol} simplifies to
2105: \begin{equation}
2106: \pmatrix{\K_{--,\vec{k}}\cr \K_{-+, \vec{k}}\cr \K^*_{--, \vec{k}}\cr
2107: \K^*_{-+, \vec{k}}}=
2108: \pmatrix{-1&B&0&B\cr
2109:          B&-1&B&0\cr
2110:          0&-B&-1&-B\cr
2111:          -B&0&-B&-1}
2112: \pmatrix{\K^*_{++, \vec{k}}\cr \K^*_{+-, \vec{k}}\cr \K_{++, \vec{k}}\cr
2113: \K_{+-, \vec{k}}},
2114: \label{bogolsimp}
2115: \end{equation}
2116: which can be inverted to
2117: \be\label{bogolsimpinv}
2118: \pmatrix{\K^*_{++, \vec{k}}\cr \K^*_{+-, \vec{k}}\cr \K_{++, \vec{k}}\cr \K_{+-,
2119: \vec{k}}}=
2120: \pmatrix{-1&-B&0&-B\cr
2121:          -B&-1&-B&0\cr
2122:          0&B&-1&B\cr
2123:          B&0&B&-1}
2124: \pmatrix{\K_{--, \vec{k}}\cr \K_{-+, \vec{k}}\cr \K^*_{--, \vec{k}}\cr
2125: \K^*_{-+, \vec{k}}}.
2126: \ee
2127: It is straightforward to check these relations using the formulas in
2128: Appendix~B.
2129: Inserting~\eq{bogolsimpinv} into~\eq{expandin}, we find the Bogolubov transformation
2130: \be\label{bogolubov}
2131: \pmatrix{a^{II\dagger}_{\vec k}\cr a^{V\dagger}_{\vec k} \cr a^{II}_{-\vec k}\cr
2132: a^{V}_{-\vec k}}=
2133: \pmatrix{-1&-B&0&B\cr
2134:          -B&-1&B&0\cr
2135:          0&-B&-1&B\cr
2136:          -B&0&B&-1}
2137: \pmatrix{a^{I\dagger}_{\vec k}\cr a^{VI\dagger}_{\vec k} \cr a^{I}_{-\vec k}\cr
2138: a^{VI}_{-\vec k}}.
2139: \ee
2140: 
2141: 
2142: Note that the Bogolubov  transformation \eq{bogolubov}
2143: mixes creation with annihilation operators. This
2144: implies particle creation at late times \cite{Craps:2002ii}.
2145: To see this, note that for each momentum $\vec{k}$
2146: \be\label{particlesII}
2147: _{in}\langle 0|a_{\vec{k}}^{II\dagger}a_{\vec{k}}^{II}|0\rangle_{in}=|B|^2={1\over\sinh^2\left(
2148: {\pi E_{\vec k}\over2Q}\right)}
2149: \ee
2150: and
2151: \be\label{particlesV}
2152: _{in}\langle 0|a_{\vec{k}}^{V\dagger}a_{\vec{k}}^{V}|0\rangle_{in}=
2153: |B|^2={1\over\sinh^2\left(
2154: {\pi E_{\vec k}\over2Q}\right)}.
2155: \ee
2156: The physical interpretation is that, at late times, the vacuum
2157: $|0\rangle_{in}$ has $1/\sinh^{2}(\pi E_{\vec{k}}/2Q)$ particles with momentum
2158: $\vec{k}$ in Region I and the same number of particles with momentum $\vec{k}$
2159: in Region V for each value of $\vec{k}$.
2160: Note that the number of particles created goes to zero%
2161: \footnote{At least for the modes \eq{pw} we are considering.
2162: The generalization of \eq{particlesII} is \cite{Craps:2002ii}
2163: \be |B|^2={\cosh^2(\pi
2164: m)\over\sinh^2\left({\pi E\over 2Q}\right)} \ee
2165: with \be m\equiv{1\over2}\left({n\over Qr_0}-{wr_0\over
2166: Q\alpha'}\right),\ee
2167: $n$ and $w$ being integers labelling momentum and winding in the
2168: $x$ direction and $E$ being the energy.
2169: }
2170: in the limit that
2171: \begin{equation}
2172: E_{\vec{k}}/Q\longrightarrow \infty.
2173: \label{eq:d1}
2174: \end{equation}
2175: We will use this result in subsection~\ref{subsec:milne}.
2176: 
2177: 
2178: 
2179: Let us now, within the context of the generalized Milne orbifold,
2180: compute the analogue of expression \eq{eq:57}, that is, the two-point
2181: correlation function
2182: \begin{equation}
2183: _{in}\langle 0|{\delta \sigma_{T}}(uv,\vec x){\delta \sigma_{T}}(uv,\vec x+\vec
2184: r)
2185: |0\rangle _{in}.
2186: \label{eq:80bis}
2187: \end{equation}
2188: %(Later on, we will put $\vec{r}=0$ as after \eq{eq:57}.)
2189: Since this function is correlated with respect to the in-vacuum, it is
2190: simplest to expand the fluctuations as in~\eq{expandin}.
2191: Using this and \eq{cancomm}, we find that
2192: \beqa
2193: &&_{in}\langle 0|{\delta \sigma_{T}}(uv,\vec x){\delta \sigma_{T}}(uv,\vec x+\vec r)
2194: |0\rangle _{in}\cr
2195: &&={1\over 2\pi^2}\int dk {|\vec{k}|}^2 \left(|\K_{++,\vec k}|^{2}+
2196: |\K_{+-,\vec k}|^{2}\right)
2197: \frac{\sin(|\vec{k}||\vec{r}|)}{|\vec{k}||\vec{r}|}.
2198: \label{eq:80tris}
2199: \eeqa
2200: As in section 1, we will always set $\vec{r}=0$ and consider
2201: \begin{equation}
2202: _{in}\langle 0|{\delta \sigma_{T}}(uv,\vec x)^{2}|0\rangle _{in}
2203: =\int{ \frac{d^{3}k}{(2\pi)^{3}} \left(|\K_{++,\vec k}|^{2}+
2204: |\K_{+-,\vec k}|^{2}\right)}.
2205: \label{eq:d2}
2206: \end{equation}
2207: Since our wavefunctions are globally defined,
2208: \eq{eq:d2} is valid for $(u,v)$ in any of the six regions of the orbifold.
2209: Of course, the explicit form of of $\K_{++,\vec k}$ and $\K_{+-,\vec k}$ change
2210: from region to region and, hence, so will the expression for the correlation
2211: function. Let us begin by calculating~\eq{eq:d2} in Region I. It follows
2212: from~\eq{eq:c1} and~\eq{eq:c7} that
2213: \be\label{eq:80trisone}
2214: _{in}\langle 0|{\delta\sigma_{T}}(t,\vec x)^{2}|0\rangle _{in}=
2215: \int \frac{d^{3}k}{(2\pi)^{3}}|\delta T_{\vec{k}}^{+}|^{2}.
2216: \ee
2217: The agreement with \eq{eq:57} illustrates that the in-vacuum
2218: defined by \eq{expandin} is indeed the correct vacuum in Region~I. As
2219: previously, for momenta $\vec{k}^{2}\gg Q^{2}$ in the distant past, the
2220: two-point function in Region I becomes
2221: \begin{equation}
2222: _{in}\langle 0|{\delta\sigma_{T}}(t,\vec x)^{2}|0\rangle _{in}=\int\frac{d^3k}{(2\pi)^{3}}
2223: \frac{1}{|\vec{k}|}\left(\frac{1}{2}+\frac{Q^{2}/2}{2|\vec{k}|^{2}}\right),
2224: \label{eq:d3}
2225: \end{equation}
2226: where we have ignored a momentum-independent factor that can be
2227: removed by a field redefinition. The same
2228: expression for the correlation function and conclusions hold in Region VI.
2229: 
2230: 
2231: Having established this, we now calculate the object of real interest, namely,
2232: the two-point function~\eq{eq:d2} in Region II. Using~\eq{bogolsimpinv} and the
2233: expressions for the wavefunctions given in the previous subsection, we find
2234: that
2235: \beqa\label{eq:80tristwo}
2236: &&_{in}\langle 0|{\delta \sigma_{T}}(uv,\vec x)^{2}|0\rangle _{in} \cr
2237: &&=\int {\frac{d^{3}k}{(2\pi)^{3}} \left((1+2|B|^{2})|\delta T_{\vec{k}}^{+}|^{2}
2238: +|B(\vec k)|^{2}\delta T_{\vec{k}}^{+2}+|B(\vec k)|^{2}
2239: \delta T_{\vec{k}}^{- 2} \right)}.
2240: \eeqa
2241: As $t\rightarrow +\infty$, this
2242: expression becomes
2243: \begin{equation}
2244: _{in}\langle0|{\delta \sigma_{T}}(uv,\vec x)^{2}|0\rangle _{in}=
2245: \int{\frac{d^{3}k}{(2\pi)^{3}}\frac{1}{|\vec{k}|}\left(\frac{1}{2}+
2246: \frac{Q^{2}/2}{2|\vec{k}|^{2}}\right)\Delta(\vec{k},t)},
2247: \label{eq:e1}
2248: \end{equation}
2249: where we have ignored a momentum-independent factor and
2250: \begin{equation}
2251: \Delta(\vec{k},t)=1+2|B|^{2}+|B|^{2}e^{-2iE_{\vec{k}}t}+|B|^{2}e^{2iE_{\vec{k}}t}.
2252: \label{eq:e2}
2253: \end{equation}
2254: This should be compared with the field theory result given in~\eq{eq:78}.
2255: Clearly, it is impossible to find
2256: coefficients $X$ and $Y$ such that \eq{eq:78} reproduces \eq{eq:e2}.
2257: This discrepancy has an interesting origin and interpretation, which will be
2258: the subject of subsection~\ref{sub:string:informationloss}.
2259: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2260: 
2261: \subsubsection{A Family of Vacuum States}\label{subsub:family}
2262: 
2263: So far, we have considered the two vacuum states of
2264: \cite{Craps:2002ii}. One of them is empty in each of the early time
2265: regions I and VI, while the other is empty in the late time regions
2266: II and V. Using these vacua, we found that there is
2267: non-trivial particle creation in Region II, described by
2268: \eq{particlesII}. These particles influence the two-point function
2269: of the corresponding field, leading to the non-trivial factor of
2270: $\Delta(\vec{k},t)$ in \eq{eq:e2}.
2271: Upon reflection, however, one is led to wonder whether the choice of vacua of
2272: \cite{Craps:2002ii} is really unique and, in particular, whether
2273: there exists alternative vacuum states for which the early time
2274: two-point function \eq{eq:80trisone} is less drastically altered
2275: upon going through the Big Crunch/Big Bang transition.
2276: 
2277: Since we are interested in an observer who starts out in Region~I and ends
2278: up in Region~II after the Big Crunch/Big Bang transition, we will continue
2279: to impose the condition that any alternative in-vacuum state should be empty in
2280: Region I. That is, the two-point function in Region~I should equal
2281: \eq{eq:80trisone}. However, we will no longer impose a similar condition
2282: in the other in-region, Region~VI, since our observer does not live there.
2283: We now construct a family of such generalized in-vacua and compute, in each
2284: case, the analogues of \eq{particlesII} and \eq{eq:e1}. These correspond
2285: to quantities our observer could, in principle, measure in Region~II.
2286: 
2287: We continue to use the expansion \eq{expand} with
2288: \be\label{K1++}
2289: \K_{1, \vec{k}}=\K_{++,\vec{k}},
2290: \ee
2291: which has positive frequency in Region~I and vanishes in Region~VI. However,
2292: we now allow a more general wavefunction for $\K_{2, \vec{k}}$, which vanishes in
2293: Region~I but does not necessarily have positive frequency in Region VI.
2294: The most general such wavefunction, for a given momentum, is an arbitrary
2295: normalized linear combination of $\K_{+-}$ and $\K_{+-}^*$ specified by
2296: \be\label{K2gen}
2297: \K_{\gamma\tilde\gamma,\vec k}=
2298: \gamma(\vec k)\K_{+-,\vec k}+\tilde\gamma(\vec k)
2299: (\K_{+-,\vec k}+\K_{+-,\vec k}^*),
2300: \ee
2301: where $\gamma(\vec k)$ and $\tilde\gamma(\vec k)$ are complex numbers satisfying
2302: \be\label{gammanorm}
2303: |\gamma(\vec k)+\tilde\gamma(\vec k)|^2-|\tilde\gamma(\vec k)|^2=1.
2304: \ee
2305: For each
2306: \be\label{gammatildegamma}
2307: \gamma(\vec k)\neq0,
2308: \ee
2309: there exists at least one $\tilde\gamma$ such that \eq{gammanorm} is
2310: satisfied. Note that $\gamma(\vec k)=1,\tilde\gamma(\vec k)=0$
2311: corresponds to the natural in-vacuum
2312: $|0\rangle_{in}$ defined using \eq{expandin}.
2313: For simplicity, we will assume that
2314: \be\label{assumegamma}
2315: \gamma(\vec k)=\gamma(-\vec k),\ \ \tilde\gamma(\vec k)=\tilde\gamma(-\vec k),
2316: \ee
2317: although more general choices would not change our conclusions.
2318: Henceforth, we will take
2319: \begin{equation}
2320: \K_{2,\vec{k}}=\K_{\gamma\tilde\gamma, \vec{k}}
2321: \label{eq:e3}
2322: \end{equation}
2323: and use the expansion
2324: \beqa
2325: &&\delta\sigma_T=\int{d^3k\over(2\pi)^{3/2}}\left(a^{I}_{\vec{k}}\K_{++,\vec
2326: k}(uv)
2327: e^{i\vec k
2328: \cdot\vec x}+a^{VI}_{\vec k}\K_{\gamma \tilde\gamma,\vec k}(uv)
2329: e^{i\vec k\cdot\vec x}\right.\cr
2330: &&\ \ \ \ \ \ \ \ \ \ \ \ \ \ \  \ \ \ \ \ \  \left.
2331: +a^{I \dagger}_{\vec{k}}\K_{++,\vec k}^*(uv)e^{-i\vec k\cdot\vec x}
2332: +a^{VI \dagger}_{\vec{k}}\K_{\gamma \tilde\gamma,\vec k}^{*}(uv)e^{-i\vec k\cdot\vec x}\right).
2333: \label{eq:e4}
2334: \eeqa
2335: 
2336: 
2337: We can now repeat the computations of subsection~\ref{subsub:quantization}
2338: for the new in-vacuum states $|0\rangle_{\gamma\tilde\gamma}$ defined by
2339: \eq{expand}, \eq{vacuumstate}, \eq{K1++},\eq{K2gen} and~\eq{gammanorm}. We continue to
2340: use the natural out-vacuum defined using \eq{expandout}. It follows from the
2341: expressions for the wavefunctions given in Appendix B and~\eq{K2gen} that
2342: \be\label{bogolsimpinvbis}
2343: \pmatrix{\K^*_{++, \vec{k}}\cr \K^*_{\gamma\tilde\gamma, \vec{k}}\cr \K_{++,
2344: \vec{k}}\cr
2345: \K_{\gamma\tilde\gamma, \vec{k}}}=
2346: \pmatrix{-1&-B&0&-B\cr
2347:          -\gamma^* B&-(\gamma^*+\tilde\gamma^*)&-\gamma^* B&-\tilde\gamma^*\cr
2348:          0&B&-1&B\cr
2349:          \gamma B&-\tilde\gamma&\gamma B&-(\gamma+\tilde\gamma)}
2350: \pmatrix{\K_{--, \vec{k}}\cr \K_{-+, \vec{k}}\cr \K^*_{--, \vec{k}}\cr
2351: \K^*_{-+, \vec{k}}},
2352: \ee
2353: where $B$ is given in~\eq{ABC}. Inserting this into~\eq{eq:e4} and comparing
2354: to~\eq{expandout}, we find the Bogolubov transformation
2355: \be\label{bogolubovbis}
2356: \pmatrix{a^{II\dagger}_{\vec k}\cr a^{V\dagger}_{\vec k} \cr a^{II}_{-\vec k}\cr
2357: a^{V}_{-\vec k}}=
2358: \pmatrix{-1&-\gamma^*B&0&\gamma B\cr
2359:          -B&-(\gamma^*+\tilde\gamma^*)&B&-\tilde\gamma\cr
2360:          0&-\gamma^*B&-1&\gamma B\cr
2361:          -B&-\tilde\gamma^*&B&-(\gamma+\tilde\gamma)}
2362: \pmatrix{a^{I\dagger}_{\vec k}\cr a^{VI\dagger}_{\vec k} \cr a^{I}_{-\vec k}\cr
2363: a^{VI}_{-\vec k}}.
2364: \ee
2365: This can be inverted to give
2366: \be\label{bogolubovtris}
2367: \pmatrix{a^{I\dagger}_{\vec k}\cr a^{VI\dagger}_{\vec k} \cr a^{I}_{-\vec k}\cr
2368: a^{VI}_{-\vec k}}=
2369: \pmatrix{-1&B&0&- B\cr
2370:         \gamma B&-(\gamma+\tilde\gamma)&-\gamma B&\tilde\gamma\cr
2371:          0&B&-1&-B\cr
2372:          \gamma^*B&\tilde\gamma^*&-\gamma^*B&-(\gamma^*+\tilde\gamma^*)}
2373: \pmatrix{a^{II\dagger}_{\vec k}\cr a^{V\dagger}_{\vec k} \cr a^{II}_{-\vec k}\cr
2374: a^{V}_{-\vec k}}.
2375: \ee
2376: Using~\eq{bogolubovbis}, we can now compute the occupation numbers
2377: \be\label{particlesIIgen}
2378: {}_{\gamma\tilde\gamma}\langle 0|a_{\vec{k}}^{II\dagger}a_{\vec{k}}^{II}|0\rangle_{\gamma\tilde\gamma}=
2379: |\gamma B|^2=
2380: {|\gamma(\vec k)|^2\over\sinh^2\left({\pi E_{\vec k}\over2Q}\right)}
2381: \ee
2382: in Region~II and
2383: \be\label{particlesVgen}
2384: {}_{\gamma\tilde\gamma}\langle 0|a_{\vec k}^{V\dagger}a_{\vec k}^V|0\rangle_{\gamma\tilde\gamma}=
2385: |B|^2+|\tilde \gamma|^2
2386: \ee
2387: in Region~V.
2388: For the reasons given at the beginning of this subsection, we are principally interested
2389: in the number of particles in Region~II. We see from \eq{particlesIIgen} that
2390: this number can be made arbitrarily small by choosing $\gamma(\vec k)$ to be close
2391: to zero. Note, however, that we cannot set $\gamma(\vec{k})=0$, since this
2392: would be inconsistent with the constraint~\eq{gammanorm}.
2393: 
2394: Next we compute
2395: \beqa\label{eq:80tristwogen}
2396: &&_{\gamma\tilde\gamma}\langle 0|{\delta \sigma_{T}}(uv,\vec x)^{2}
2397: |0\rangle _{\gamma\tilde\gamma}\cr
2398: &&=\int \frac{d^{3}k}{(2\pi)^{3}} \left(|\K_{++,\vec k}|^{2}+ |
2399: \K_{\gamma,\tilde\gamma,\vec k}|^{2}\right).
2400: \eeqa
2401: In Region~I, $\K_{+-, \vec{k}}$ vanishes. It follows that $\K_{\gamma
2402: \tilde\gamma, \vec{k}}$ also vanishes and, hence,
2403: \eq{eq:80tristwogen} reduces to the familiar
2404: result \eq{eq:80trisone}.
2405: From \eq{KII} in Appendix~B, we see that $\K_{+-, \vec{k}}$ is purely imaginary
2406: in Region~II.
2407: In fact, this is also the case in the other regions that touch $uv=0$. Therefore,
2408: $\K_{\gamma\tilde\gamma, \vec{k}}$ reduces to $\gamma\K_{+-, \vec{k}}$
2409: in those regions. Using this,
2410: \eq{eq:80tristwogen} takes the following form in Region~II,
2411: \beqa\label{eq:80tristwogenbis}
2412: &&_{\gamma\tilde\gamma}\langle 0|{\delta \sigma_{T}}(uv,\vec x)^{2}
2413: |0\rangle _{\gamma\tilde\gamma}\cr
2414: &&=\int \frac{d^{3}k}{(2\pi)^{3}} \left((1+2|\gamma B|^{2})\,
2415: |\delta T_{\vec{k}}^{+}|^{2}
2416: +|\gamma(\vec k)B(\vec k)|^2\,
2417: \delta T_{\vec{k}}^{+2}+ |\gamma(\vec k)B(\vec k)|^2\,\delta
2418: T_{\vec{k}}^{-2} \right).\cr &&
2419: \eeqa
2420: In the far future where $t\rightarrow+\infty$, and for $\vec k^2\gg Q^2$,
2421: this expression becomes
2422: \begin{equation}
2423: _{\gamma\tilde\gamma}\langle 0|{\delta \sigma_{T}}(uv,\vec x)^{2}
2424: |0\rangle _{\gamma\tilde\gamma}=\int
2425: \frac{d^{3}k}{(2\pi)^{3}}\frac{1}{|\vec{k}|}
2426: \left(\frac{1}{2}+\frac{Q^{2}/2}{2|\vec{k}|^{2}}\right)\Delta(\vec{k},t),
2427: \label{eq:e5}
2428: \end{equation}
2429: where we have dropped a momentum-independent scale factor and
2430: \begin{equation}
2431: \Delta(\vec{k},t)=1+2|\gamma B|^{2}+|\gamma B|^{2}e^{-2iE_{\vec{k}}t}+
2432: |\gamma B|^{2}e^{2iE_{\vec{k}}t}.
2433: \label{eq:e6}
2434: \end{equation}
2435: In the limit $\gamma (\vec k)\to 0$, this approaches the early time result
2436: \eq{eq:d3}. Therefore, for this limiting choice of in-vacuum, an observer in
2437: Region~II
2438: would say that the fluctuations in Region~I went through the Big Crunch/Big
2439: Bang unchanged.
2440: For this special limiting case, the result \eq{eq:e6} can be reformulated
2441: in the framework of subsection~\ref{sub:BBBC}. It corresponds to the Bogolubov
2442: coefficients $X$ and $Y$ with the properties
2443: \be\label{XY}
2444: |X|=1,\ \ Y=0\ \ \ \ \ {\rm in\ the\ limit}\ \gamma(\vec k)\to0.
2445: \ee
2446: However, this limiting case is the only one for which the framework of
2447: subsection~\ref{sub:BBBC} can be made to reproduce \eq{eq:e6}. This
2448: apparent discrepancy is the subject of the next subsection.
2449: 
2450: 
2451: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2452: 
2453: \subsection{Information Loss}\label{sub:string:informationloss}
2454: 
2455: It is useful to use the Bogolubov transformation~\eq{bogolubovtris} to
2456: explicitly write the relation between the in-vacua
2457: $|0\rangle_{\gamma\tilde\gamma}$ and the natural out-vacuum $|0\rangle_{out}$.
2458: We find that (remember \eq{assumegamma})
2459: \be\label{zerogammaout}
2460: |0\rangle_{\gamma\tilde\gamma}=N_{\gamma\tilde\gamma}
2461: \exp\left\{\int d^3k\left(\theta a^{II^\dagger}_{\vec{k}}a^{II^\dagger}_{-\vec{k}}
2462: +\lambda a^{V\dagger}_{\vec{k}}a^{V\dagger}_{-\vec{k}}
2463: +\mu a^{II\dagger}_{\vec{k}}a^{V\dagger}_{-\vec{k}}\right)\right\}\,
2464: |0\rangle_{out},
2465: \ee
2466: where
2467: \beqa\label{thetalambda}
2468: \theta&=&-{\gamma^* B^2\over2(\gamma^*(1-B^2)+\tilde\gamma^*)},\cr
2469: \lambda&=&{\tilde\gamma^*-\gamma^*B^2\over2(\gamma^*(1-B^2)+\tilde\gamma^*)},\cr
2470: \mu&=&{\gamma^*B\over\gamma^*(1-B^2)+\tilde\gamma^*}.
2471: \eeqa
2472: To check this, use \eq{bogolubovtris} to verify that $|0\rangle_{\gamma\tilde\gamma}$
2473: defined by \eq{zerogammaout} is indeed annihilated by $a^{I}_{\vec{k}}$ and
2474: $a^{VI}_{\vec{k}}$ given that
2475: $|0\rangle_{out}$ is annihilated by $a^{II}_{\vec{k}}$ and $a^{V}_{\vec{k}}$.
2476: $N_{\gamma\tilde\gamma}$
2477: is a normalization constant, which we will determine for a special case. It is useful
2478: to keep in mind that $B$ given in~\eq{ABC} is purely imaginary.
2479: 
2480: For the natural in-vacuum $|0\rangle_{in}$ defined in \eq{expandin}, which corresponds to
2481: choosing $\gamma(\vec k)=1,\tilde\gamma(\vec k)=0$, the coefficients
2482: in~\eq{thetalambda} become
2483: \beqa\label{thetaone}
2484: \theta&=&-{B^2\over2(1-B^2)},\cr
2485: \lambda&=&-{B^2\over2(1-B^2)},\cr
2486: \mu&=&{B\over1-B^2}.
2487: \eeqa
2488: In this case, we find
2489: \be\label{None}
2490: N_{10}=\prod_{\vec k}\left(1-B^2(\vec k)\right)^{-1/2}\equiv
2491: \exp\left\{ V\int{d^3k\over(2\pi)^3}\log\left(1-B^2(\vec k)\right)^{-1/2}\right\},
2492: \ee
2493: with $V$ the volume of space. For strictly infinity $V$, $N_{10}$ vanishes, so we will
2494: assume $V$ to be large but finite instead.
2495: Then \eq{None} can be derived as follows. Define the unitary matrix
2496: \be\label{S}
2497: S=\exp\left\{\int d^3k\left(B(a^{II}_{\vec{k}}a^{V}_{-\vec{k}}
2498: +a^{II\dagger}_{\vec{k}}
2499: a^{V\dagger}_{-\vec{k}}-a^{II}_{\vec{k}}a^{V\dagger}_{\vec{k}}
2500: -a^{II\dagger}_{\vec{k}} a^{V}_{\vec{k}})\right)\right\}.
2501: \ee
2502: Using the formula
2503: \be
2504: e^{-A}Be^A=B+[B,A]+{1\over2!}[[B,A],A]+\cdots,
2505: \ee
2506: it is straightforward to verify that
2507: \beqa\label{SS}
2508: a^{I}_{\vec{k}}=-Sa^{II}_{\vec{k}}S^{-1}, \qquad
2509: a^{VI}_{\vec{k}}=-Sa^{V}_{\vec{k}}S^{-1}.
2510: \eeqa
2511: This implies that, up to a phase
2512: \be\label{Sinout}
2513: |0\rangle_{in}=S|0\rangle_{out}.
2514: \ee
2515: The normalization factor \eq{None} can then be obtained by
2516: determining the coefficient of $|0\rangle_{out}$ upon expanding the right
2517: hand side of \eq{Sinout}.\footnote{In doing this, it is convenient to work with
2518: rescaled oscillators $\alpha_{\vec k}\equiv\left((2\pi)^3/V\right)^{1/2}a_{\vec k}$,
2519: which satisfy the canonical commutation relation $[\alpha_{\vec k},\alpha^\dagger_{\vec k}]=1$.}
2520: 
2521: As another special case, consider the limiting values $\gamma(\vec k)\to0$,
2522: $\tilde\gamma(\vec k)\to\infty$. This was the limit that turned off particle creation
2523: in Region~II, as we discussed at the end of subsection~\ref{subsub:family}.
2524: This corresponds to
2525: \beqa\label{thetazero}
2526: \theta&\to&0,\cr
2527: \lambda&\to&{1\over2},\cr
2528: \mu&\to&0,
2529: \eeqa
2530: and
2531: \be\label{zerogammaoutbis}
2532: |0\rangle_{0\infty}\to N_{0\infty}
2533: \exp\left\{\int d^3k\left({1\over2}(a^{V\dagger}_{\vec{k}}a^{V\dagger}_{-\vec{k}})\right)\right\}\,
2534: |0\rangle_{\rm out}.
2535: \ee
2536: However, it turns out that in the limit \eq{thetazero} the normalization factor
2537: $N_{0,\infty}$ vanishes, even for finite $V$.
2538: 
2539: We have not computed the normalization factor for generic values of $\gamma$ and
2540: $\tilde\gamma$, although it could, in principle, be done in the same way. For instance,
2541: if $\gamma$ and $\tilde\gamma$ are both real and positive, the generalization of \eq{S} is
2542: \beqa\label{Sbis}
2543: S&=&\exp\left\{\int d^3k\left(-\rho(a^{II}_{\vec{k}}a^{V}_{-\vec{k}}+
2544: a^{II\dagger}_{\vec{k}}
2545: a^{V\dagger}_{-\vec{k}}-a^{II}_{\vec{k}}a^{V\dagger}_{\vec{k}}
2546: -a^{II\dagger}_{\vec{k}} a^{V}_{\vec{k}})\right.\right.\cr
2547: &&\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \left.\left.+\mu(a^{V}_{\vec{k}}a^{V}_{-\vec{k}}
2548: -a^{V\dagger}_{\vec{k}}a^{V\dagger}_{-\vec{k}})\right)\right\},
2549: \eeqa
2550: with
2551: \beqa\label{sigmatau}
2552: \cosh(2\mu)&=&\gamma+\tilde\gamma, \cr
2553: \sinh(2\mu)&=&-\tilde\gamma, \cr
2554: \rho {e^{2\mu}-1\over2\mu}&=&-\gamma B.
2555: \eeqa
2556: If $\gamma$ and $\tilde \gamma$ do not have the same phase, the exponent of $S$
2557: also involves $a^{II}_{\vec{k}}a^{II}_{-\vec{k}}$ and
2558: $a^{II\dagger}_{\vec k}a^{II\dagger}_{-\vec k}$.
2559: 
2560: The Hilbert space of states ${\cal H}$ we have been using can be viewed as
2561: the tensor product
2562: \be\label{tensor}
2563: {\cal H}=\H^{II}\otimes\H^V
2564: \ee
2565: of a Hilbert space $\H^{II}$ associated with $a^{II}_{\vec{k}}$ and
2566: $a^{II\dagger}_{\vec{k}}$ and a Hilbert space $\H^V$ associated with
2567: $a^{V}_{\vec{k}}$
2568: and $a^{V\dagger}_{\vec{k}}$. These two Hilbert spaces can be thought of as
2569: associated with the out-regions~II and~V respectively. Note that $\H$ could equally well be
2570: written as a tensor product
2571: of Hilbert spaces associated with in-regions~I and~VI.
2572: The out-vacuum can, accordingly, be expressed as
2573: \be\label{outvac}
2574: |0\rangle_{out}=|0\rangle_{II}\otimes|0\rangle_V.
2575: \ee
2576: It is clear
2577: that each in-vacuum defined by \eq{zerogammaout} is a pure state in this
2578: tensor product Hilbert space. With a pure state, one can
2579: associate a trivial density matrix
2580: \be\label{density}
2581: \rho_{\gamma\tilde\gamma}=
2582: |0\rangle_{\gamma\tilde\gamma}\,{}_{\gamma\tilde\gamma}\langle0|.
2583: \ee
2584: When one is only interested in computing correlation functions in
2585: Region~II, which would be the case for a physicist living in
2586: that region and trying to predict the result of experiments done
2587: there, it is convenient (and equivalent) to use the density matrix obtained by
2588: tracing \eq{density} over $\H^V$. This is given by
2589: \be\label{densitytwo}
2590: \rho^{II}_{\gamma\tilde\gamma}=\sum_i{}_V\langle i|\rho_{\gamma\tilde\gamma}
2591: |i\rangle_V,
2592: \ee
2593: where $\{|i\rangle_{V}\}$ is an orthonormal basis of $\H^V$. Note that
2594: $|0\rangle_{\gamma\tilde\gamma}\in{\cal
2595: H}=\H^{II}\otimes\H^V$, whereas $|i\rangle_V\in\H^V$. Therefore,
2596: \eq{densitytwo} is indeed a density matrix in Region~II, that is,
2597: an element of $\H^{II*}\otimes\H^{II}$. Now, if $\mu$ in
2598: \eq{zerogammaout} is nonzero (as it is, except in the limit
2599: $\gamma\to0$), $|0\rangle_{\gamma\tilde\gamma}$ is an entangled state in
2600: ${\cal H}=\H^{II}\otimes\H^V$. When one traces the density matrix
2601: of an entangled pure state over $\H^V$, one obtains a non-trivial
2602: density matrix \eq{densitytwo} in Region~II. Therefore, an observer in
2603: Region~II with no access to information about Region~V finds
2604: himself in a mixed state, that is, a state with entropy. This entropy
2605: is called ``entanglement entropy'' and reflects the ignorance
2606: about the correlated Region~V. In the limit $\gamma\to0$, one again obtains a
2607: trivial density matrix
2608: \be\label{densityzero}
2609: \rho^{II}_{0\infty}=|0\rangle_{II}\,{}_{II}\langle0|,
2610: \ee
2611: which is consistent with \eq{XY}.
2612: 
2613: This explains the discrepancy noted at the end of
2614: subsection~\ref{subsub:family}. In subsection~\ref{sub:BBBC}, we
2615: assumed unitary evolution from Region~I to Region~II, that is, a
2616: pure state in Region~I evolving into a pure (squeezed) state in
2617: Region~II. This assumption generically turns out to be
2618: inconsistent with what we find in the full string theory
2619: background, which includes additional regions. We find that a pure
2620: in-vacuum state evolves to a pure state in the tensor product
2621: Hilbert space $\H$, but that this state generically contains
2622: correlations between regions~II and~V. As a consequence, in a
2623: description where Region~V is ignored, such as the one
2624: appropriate for an observer in Region~II, this state is a mixed
2625: state. Only in the limit $\gamma\to0$ does the in-vacuum evolve to
2626: a pure state in Region~II. As explained in
2627: subsection~\ref{subsub:family}, the limit $\gamma\to0$ can,
2628: strictly speaking, not be reached. The more accurate statement is that
2629: the entropy of the state as described in Region~II can be made
2630: arbitrarily small by taking $\gamma$ to be arbitrarily close to zero.
2631: 
2632: The above discussion is a generalization of thermofield dynamics as
2633: recently applied to the eternal BTZ black hole, where the thermal
2634: state of a scalar field outside the black hole is obtained from an entangled
2635: state in a tensor product of two Hilbert spaces, the second Hilbert space
2636: being associated with an additional asymptotic region behind the horizon
2637: of the black hole \cite{Maldacena:2001kr}. Also see
2638: \cite{BTZ} and references therein.
2639: 
2640: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2641: 
2642: \subsection{The Milne Orbifold Limit}\label{subsec:milne}
2643: In the limit that
2644: \be\label{Milnelim}
2645: uv\to0,
2646: \ee
2647: the metric in~\eq{backgroundIIA} reduces to the Milne orbifold metric \eq{Milne}
2648: \be\label{Milnemetric}
2649: ds^2={1\over Q^2}dudv+\sum^{8}_{I=1}(dx^I)^2
2650: \ee
2651: subject to the identification \eq{uvid},
2652: \be\label{identorb}
2653: (u,v)\sim(ue^{-2\pi QR},ve^{2\pi QR}).
2654: \ee
2655: In addition, it follows from~\eq{backgroundIIA} that, in this limit,
2656: \begin{equation}
2657: \Phi\longrightarrow \Phi_{0}.
2658: \label{eq:h1}
2659: \end{equation}
2660: Note that the theory we are explicitly discussing in this
2661: paper, \eq{backgroundIIA}, differs in the number of
2662: compactified dimensions from the Milne orbifold
2663: \eq{Milne}. Specifically, the generalized Milne orbifold has $(d+1)$-dimensions,
2664: whereas the Milne space has ten dimensions. How, then, can we claim that
2665: metric \eq{backgroundIIA} reduces near the singularity to \eq{Milne}? The
2666: naive answer is that this limit does not involve the internal spatial
2667: dimensions which, therefore, are irrelevant. A more precise way of
2668: understanding the Milne limit is to recognize, as we mention in the
2669: Introduction, that two flat spatial directions of the generalized Milne
2670: orbifold can be replaced by a ``cigar'' geometry \cite{Witten:1991yr}
2671: without changing any of the
2672: conclusion of this paper. In this case, we have a critical, ten-dimensional
2673: string theory where $Q$ is a free parameter. This cigar geometry has a
2674: well-defined smooth limit to the Milne orbifold. Keeping this in mind, we will
2675: ignore the number of compact dimensions in this subsection.
2676: 
2677: If in the limit \eq{Milnelim} we keep the ratio
2678: \be\label{fixed}
2679: \frac{{\vec{k}}^{2}uv}{Q^2} \quad {\rm fixed},
2680: \ee
2681: then the fluctuation equation \eq{eq:37} on the generalized Milne
2682: orbifold reduces to
2683: \be\label{eommilne}
2684: \delta T''+{1\over t}\delta T'+\vec k^2\delta T=0.
2685: \ee
2686: Were we to consider fields with non-vanishing momentum along the $x$
2687: circle as well, we would also be required to keep the product
2688: \begin{equation}
2689: Qr_{0} \quad {\rm fixed}.
2690: \label{eq:g1}
2691: \end{equation}
2692: Expression~\eq{eommilne} is precisely the fluctuation equation for a
2693: scalar mode in the Milne orbifold. The limit defined
2694: by~\eq{Milnelim},~\eq{fixed} and~\eq{eq:g1}
2695: takes the generalized Milne orbifold to the Milne orbifold.
2696: 
2697: 
2698: It is instructive to see what our
2699: global wavefunctions $\K_{\pm\pm, \vec{k}}$ reduce to in this limit. In
2700: Appendix~C, we show that $\K_{++, \vec{k}}$ reduces to the standard
2701: wavefunctions defining the adiabatic vacuum of the Milne orbifold,
2702: that is, superpositions of positive frequency plane waves in Minkowski
2703: space. They are Hankel functions in regions~I and~II and
2704: modified Bessel functions that decay asymptotically in regions~III
2705: and~IV, the expressions in
2706: the four regions being related by analytic continuation. Near the Big
2707: Crunch/Big Bang singularity, these wavefunctions diverge
2708: logarithmically. On the other
2709: hand, we find that
2710: \begin{equation}
2711: \K_{+-, \vec{k}}\longrightarrow 0
2712: \label{eq:g2}
2713: \end{equation}
2714: everywhere in the Milne limit. More precisely, it
2715: is always zero in Region I, is proportional to a finite Bessel function in
2716: Region II and corresponds to an asymptotically growing modified Bessel function in
2717: regions~III and~IV. Near the
2718: Big Crunch/Big Bang singularity, $\K_{+-, \vec{k}}$ approachs constants in
2719: all regions. However, its overall normalization factor vanishes, suppressing
2720: the wavefunction completely. It then follows from~\eq{K2gen} that, in this limit,
2721: \begin{equation}
2722: \K_{\gamma\tilde \gamma, \vec{k}}\longrightarrow 0.
2723: \label{eq:g3}
2724: \end{equation}
2725: Inserting these results into expression~\eq{eq:80tristwogen} for the scalar
2726: two-point function, we find that, in the Milne limit, for any choice of
2727: in-vacuum
2728: \beqa\label{eq:f1}
2729: _{\gamma\tilde\gamma}\langle 0|{\delta \sigma_{T}}(uv,\vec x)^{2}
2730: |0\rangle _{\gamma\tilde\gamma}
2731: =\int \frac{d^{3}k}{(2\pi)^{3}} |\K_{++,\vec k}|^{2},
2732: \eeqa
2733: which is valid in all regions. This is the standard result for
2734: the two-point function in the adiabatic
2735: vacuum of the Milne orbifold. It is useful to evaluate this correlation
2736: function in regions I and II. It is straightforward to do this in Region I
2737: using~\eq{eq:c1}. To evaluate~\eq{eq:f1} in Region II, first note
2738: from~\eq{Milnelim}
2739: and~\eq{fixed} that, in the Milne limit,
2740: \begin{equation}
2741: E_{\vec k}/Q=
2742: \sqrt{\vec k^2-Q^2}/Q\to\infty
2743: \label{eq:f2}
2744: \end{equation}
2745: and, hence, that the Bogolubov coefficient given in~\eq{ABC} satisfies
2746: \begin{equation}
2747: B\longrightarrow 0.
2748: \label{eq:f3}
2749: \end{equation}
2750: One can then compute the correlation function in Region II using~\eq{eq:c9}
2751: and the Bogolubov transformation~\eq{bogolsimpinvbis}. The result is that
2752: in both regions I and II of the Milne orbifold,
2753: \be\label{eq:f4}
2754: _{in}\langle 0|{\delta\sigma_{T}}(t,\vec x)^{2}|0\rangle _{in}=
2755: \int \frac{d^{3}k}{(2\pi)^{3}}|\delta T_{\vec{k}}^{+}|^{2}.
2756: \end{equation}
2757: It follows from our previous discussions that, in the distant past and distant
2758: future, the spectrum is of the form of Minkowski fluctuations plus a subdominant scale invariant contribution and, therefore,
2759: \begin{equation}
2760: \Delta(\vec{k},t)=1.
2761: \label{eq:g5}
2762: \end{equation}
2763: That is, in the Milne orbifold, the spectrum is unchanged by
2764: passing through the Big Bang/Big Crunch singularity. Note that this quantum
2765: process can be described by the pure Region I and Region II four-dimensional
2766: effective theory with the Bogolubov coefficients
2767: \begin{equation}
2768: |X|=1, \qquad Y=0.
2769: \label{eq:g6}
2770: \end{equation}
2771: 
2772: Finally, we see that from~\eq{particlesII} and~\eq{eq:f3} that, in the Milne
2773: limit,
2774: \begin{equation}
2775: _{in}\langle 0|a_{\vec{k}}^{II\dagger}a_{\vec{k}}^{II}|0\rangle_{in}
2776: \longrightarrow 0.
2777: \label{eq:g7}
2778: \end{equation}
2779: That is, particle creation is turned off.  We should mention that we have
2780: assumed that $\gamma$ is kept fixed during the limiting proceedure.
2781: It is clear that there would
2782: be particle creation if, instead, one kept $|\gamma B|$ fixed during this limit.
2783: 
2784: 
2785: 
2786: 
2787: 
2788: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2789: 
2790: \subsection{A Note on Backreaction}\label{sub:string:backreaction}
2791: 
2792: As we have emphasized at the end of the Introduction, the
2793: string theory background \eq{backgroundIIA} has been argued to be
2794: unstable to gravitational backreaction. Hence, reliable computations
2795: should take this backreaction into
2796: account. This seems to be out of reach at present and we will not attempt to resolve
2797: the associated deep puzzles in string theory here. However, we would like to
2798: offer a few remarks from the point of view taken in this paper.
2799: 
2800: First, it is worth translating the issue into the
2801: four-dimensional language of section~\ref{sec:fourd}.
2802: Looking at the energy density \eq{eq:13} of the
2803: four-dimensional matter fields, one might be tempted to conclude
2804: that quantum mechanical particles should not lead to large
2805: backreaction. The reason is that the energy density of the classical matter
2806: fields, since it is dominated by scalar kinetic energy,
2807: scales like $|t|^{-3}$ near the Big Bang/Big Crunch singularity at $t=0$.
2808: The wavefunction of a particle diverges at
2809: most logarithmically near $t=0$, so, according to
2810: \eq{eq:21}, its energy density cannot diverge more quickly than
2811: that of the background matter fields. It would seem, therefore, that a
2812: small fluctuation remains small compared to the background. So
2813: where is the large backreaction?
2814: 
2815: The point is that the framework of section~\ref{sec:fourd} is
2816: hard to use for computations near the bounce, since the geometry
2817: is singular there. The right framework to use is
2818: the one of section~\ref{sec:string}.
2819: In these variables, there is not much energy in the classical
2820: matter fields near the bounce, where spacetime is locally flat with
2821: a slowly varying dilaton. The reflection of this in
2822: section~\ref{sec:fourd} is that the sum of the energy densities in
2823: the matter and gravitational fields diverges only like $1/|t|$, as
2824: one can see from \eq{eq:13} and \eq{eq:30} (divided by $\kappa_4^2$).
2825: This sum is dominated by the potential energy of the
2826: four-dimensional dilaton $\phi$ which, in the string theory
2827: framework, arises from the cosmological constant. Therefore, the criterion
2828: for our computations to be unstable to backreaction is that
2829: the energy density of small fluctuations grows faster than $1/|t|$
2830: near the singularity. Unfortunately, this is the case for generic fluctuations, whose
2831: wavefunctions diverge logarithmically at the bounce. Hence, we
2832: recover the familiar large backreaction problem.
2833: 
2834: It is interesting to note that there is an important class of
2835: fluctuations that do not give rise to large backreaction
2836: \cite{Berkooz:2002je}. Of the modes considered in this paper, those
2837: described by $\K_{+-, \vec{k}}$ are of this type, as can be seen using
2838: Appendix~B. Their energy density actually vanishes at the Big Bang/Big
2839: Crunch singularity.%
2840: \footnote{This is not true for the $\K_{+-, \vec{k}}$ of the more generic
2841: string modes considered in \cite{Craps:2002ii}. For example,
2842: $\K_{+-, \vec{k}}$ does not
2843: correspond to a chiral wavefunction near the singularity if it has
2844: non-zero momentum along the $x$ circle.}
2845: 
2846: 
2847: The issue of backreaction is crucial from various points of view,
2848: string theoretic as well as cosmological. A source of criticism of the
2849: simple orbifold-like cosmological singularities studied in the string
2850: theory literature is that they are unstable against even a single particle
2851: being added to the system before the singularity \cite{Horowitz:2002mw}. Such a
2852: particle would change the conical singularity into a genuine curvature
2853: singularity. In our model, this is indeed the case for particles with generic
2854: wavefunctions, as we have just discussed. Now suppose that, however remote the
2855: possibility, no such small fluctuation appeared and the Universe made
2856: it through the Crunch singularity. Then, we have shown that the Big Crunch/Big Bang creates
2857: a collection of particles with occupation numbers \eq{particlesIIgen} in
2858: Region~II.
2859: One can wonder how these particles backreact on the geometry. First, will the
2860: energy density of these particles overwhelm the classical energy density very close
2861: to the Big Crunch/Big Bang singularity? The answer is no, as can be seen from
2862: \eq{eq:80tristwogen}. The
2863: wavefunctions of the created particles are precisely proportional to $\K_{+-,
2864: \vec{k}}$ near
2865: the singularity, so their energy density vanishes there.%
2866: \footnote{Again, this would be different for the modes we ignored in this paper. In order
2867: to turn off their backreaction in Region~II, one would have to work in a $\gamma\to0$
2868: limit.}
2869: Second, it is tempting to speculate that, in some more realistic version of our model,
2870: this particle creation might play a role as a reheating mechanism. In fact, reheating by
2871: gravitational particle creation has been discussed before. See, for instance,
2872: \cite{grav}.
2873: \medskip
2874: 
2875: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2876: \section*{Acknowledgments}
2877: We would like to thank C.~Armendariz-Picon, J.~Khoury and especially D.~Kutasov
2878: for very useful discussions. We are also grateful to the organizers and participants of
2879: the workshop ``Cosmological Perturbations on the Brane'', Cambridge, UK, for many
2880: stimulating conversations. In addition, B.~C.\ would like to thank the High Energy
2881: Theory Group at Penn for hospitality.
2882: The work of B.~C.\ is supported in part
2883: by DOE grant DE-FG02-90ER40560 and by NSF grant PHY-9901194. B.~A.~O.\ is
2884: supported in part by DOE under contract No.\ DE-AC02-76-ER-03071.
2885: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2886: 
2887: \section*{Appendix A: Global Wavefunctions from String Theory}
2888: 
2889: In this Appendix, we will outline the procedure for constructing the globally
2890: defined wavefunctions on both the Milne orbifold and the generalized Milne
2891: orbifold discussed in this paper.
2892: 
2893: \paragraph{Milne Orbifold:}
2894: We first review how globally defined wavefunctions are constructed on the
2895: Milne orbifold $\Rbar^{1,1}/\Zbar\times\Rbar^8$, an orbifold of
2896: Minkowski space.
2897: \begin{enumerate}
2898: \item
2899: Solve the wave equation
2900: \be\label{waveeq}
2901: (\nabla^2+m^2)\phi=0
2902: \ee
2903: on the covering space $\Rbar^{1,9}$. A smooth basis of solutions is
2904: given by the plane waves
2905: \be\label{planewave}
2906: \{e^{i\vec p\cdot\vec X}e^{i(p^+X^-+p^-X^+)}|2p^+p^--\vec p^2=m^2\}.
2907: \ee
2908: \item
2909: Choose an alternative basis of solutions to \eq{waveeq}, consisting of
2910: continuous
2911: superpositions of the smooth solutions \eq{planewave}, such that the
2912: orbifold generator $X^{\pm}\mapsto e^{\pm 2\pi}X^\pm$ is diagonal. Such a
2913: basis is given by \cite{Nekrasov:2002kf} (see also \cite{Berkooz:2002je})
2914: \be\label{Nekrwavef}
2915: \phi_{l,\vec p}=e^{i\vec p\cdot\vec X}\int_{-\infty}^\infty dw
2916: e^{i(p^+X^-e^{-w}+p^-X^+e^w)}e^{ilw}
2917: \ee
2918: with $l\in
2919: \Rbar$. The orbifold generator acts by multiplication by
2920: $e^{-2\pi il}$. The wavefunctions \eq{Nekrwavef} are generically not
2921: smooth near the light-cone $X^+X^-=0$, which divides the covering
2922: space in four regions. When studying string theory on Minkowski space,
2923: one could in principle use these singular wavefunctions but, since
2924: there exist smooth wavefunctions \eq{planewave} which can be written
2925: as continuous superpositions of \eq{Nekrwavef}, it is preferable to
2926: work with the latter.
2927: \item
2928: Restrict the wavefunctions \eq{Nekrwavef} to the orbifold invariant
2929: ones, that is, to those with $l\in\Zbar$. Using coordinates $t,x$ defined by
2930: $X^\pm=te^{\pm x}/\sqrt2$, in terms of which the orbifold generator
2931: acts as $x\mapsto x+2\pi$, it is clear that $l$ is the momentum in the
2932: $x$ direction. Therefore, the condition $l\in\Zbar$ is the usual momentum
2933: quantization. Hence, we end up with wavefunctions \eq{Nekrwavef} on the
2934: orbifold space. They are not smooth at the singularity $X^+X^-=0$, yet
2935: they are globally defined. Note that, because of the quantization of
2936: $l$, it is not possible to use a basis of smooth wavefunctions on the
2937: orbifold space.
2938: \item
2939: Introduce twisted sectors (winding modes). We will not discuss those
2940: in this paper; see \cite{Berkooz:2003bs} for a very recent discussion.
2941: \end{enumerate}
2942: \paragraph{Generalized Milne Orbifold:}
2943: We now review the prescription of \cite{Elitzur:2002rt,Craps:2002ii} for
2944: constructing global wavefunctions on the generalized Milne
2945: orbifold, which in string theory corresponds to a $\Zbar$ orbifold of a
2946: coset conformal field theory at negative level \cite{Craps:2002ii},
2947: \be\label{genmil}
2948: {PSL(2,\Rbar)_{k<0}\over U(1)}/\Zbar\  \times\Rbar^{d-1}.
2949: \ee
2950: \begin{enumerate}
2951: \item
2952: Start with wavefunctions on $AdS_3$, that is, the well-known
2953: eigenfunctions of the Laplacian on $AdS_3$
2954: which coincides with the $SL(2,\Rbar)$ group manifold. $SL(2,\Rbar)$ is the
2955: group of $2\times2$ matrices $g$ with unit determinant,
2956: \be
2957: g=\pmatrix{a&b\cr c&d},\ \ \ ad-bc=1.
2958: \ee
2959: $PSL(2,\Rbar)$ is obtained from this by identifying $g$ with $-g$.
2960: Wavefunctions on $PSL(2,\Rbar)$ should be consistent with this
2961: identification. The wavefunctions on a group manifold coincide with
2962: the matrix elements in the different representations in the group. Therefore,
2963: we can write
2964: \be\label{statesSL}
2965: \phi_{j,\alpha,\beta}(g)=\langle j,\alpha|g|j,\beta\rangle,
2966: \ee
2967: where $j$ labels a $PSL(2,\Rbar)$ representation and $\alpha,\beta$
2968: label states in the representation. In this paper, we will only
2969: consider the principal continuous representations, with $j=-1/2+is$
2970: and $s$ real. They correspond to taking $\vec k^2>Q^2$. A choice of basis of
2971: wavefunctions now corresponds to a choice of states $\alpha,\beta$ in
2972: the representation $j$. By choosing an appropriate basis, one can
2973: choose a smooth basis of wavefunctions on $PSL(2,\Rbar)$. Note that
2974: the group manifold $PSL(2,\Rbar)$ admits a $PSL(2,\Rbar)_L\times
2975: PSL(2,\Rbar)_R$ symmetry, the two factors acting on $g$ by left and
2976: right multiplication respectively. It is clear from \eq{statesSL}
2977: that the action of $PSL(2,\Rbar)_L$ is determined by the state
2978: $\langle j,\alpha|$, while the action of $PSL(2,\Rbar)_R$ is
2979: determined by $|j,\beta\rangle$. The Laplacian on the $PSL(2,\Rbar)$
2980: group manifold is the Casimir operator of $PSL(2,\Rbar)$ (acting either from
2981: the left or from the right). Its eigenvalue is $j(j+1)$.
2982: \item
2983: Choose a basis of wavefunctions on $PSL(2,\Rbar)$ such that the
2984: $U(1)_L\times U(1)_R$ subgroup, with both factors generated by
2985: $\sigma_3$, is diagonalized. This amounts to choosing the basis
2986: vectors of the representation $j$ such that $\sigma_3$ is diagonal. It
2987: turns out \cite{Vilenkin} that, for the representations we are
2988: considering, there are two states for each $\sigma_3$ eigenvalue
2989: $m$.%
2990: \footnote{To be precise,
2991: $e^{\alpha\sigma_3}|j,m,\pm\rangle=e^{2im\alpha}|j,m,\pm\rangle$.}
2992: Thus, we find four independent wavefunctions for each $j,m,\bar
2993: m$
2994: \be\label{KSL}
2995: K_{\pm\pm}(j,m,\bar m;g)=\langle j,m,\pm|g|j,\bar m,\pm\rangle.
2996: \ee
2997: One finds that the $PSL(2,\Rbar)$ group manifold splits into six
2998: regions where these wavefunctions are smooth. Generically, the
2999: wavefunctions are not smooth at the boundaries separating these
3000: regions. However, they are still globally defined.
3001: When studying $PSL(2,\Rbar)$ by itself, one would usually prefer to
3002: work with a basis of smooth wavefunctions.
3003: \item
3004: The coset conformal field theory $PSL(2,\Rbar)_{k<0}/U(1)$, where $k$
3005: is the level of the $PSL(2,\Rbar)$ affine Lie algebra and $U(1)$ acts
3006: as $g\mapsto e^{\alpha\sigma}ge^{\alpha\sigma}$, describes the
3007: generalized Milne spacetime \eq{backgroundIIA} without the identification
3008: \eq{uvid} \cite{Kounnas:1992wc,Craps:2002ii}.
3009: Its globally defined wavefunctions are obtained by
3010: restricting to those wavefunctions \eq{KSL} that are invariant under
3011: $U(1)$, that is, to those satisfying $m=-\bar m$
3012: \cite{Dijkgraaf:1991ba,Craps:2002ii}.
3013: \item
3014: The $\Zbar$ orbifold group is another (discrete) subgroup of
3015: $U(1)_L\times U(1)_R$. The globally defined
3016: orbifold wavefunctions are obtained by
3017: imposing a quantization condition on $m$ (momentum quantization) and
3018: by introducing twisted sectors (winding modes). We refer to
3019: \cite{Craps:2002ii} for details. In the present paper, we are
3020: interested in modes with no momentum or winding along the Milne circle.
3021: \end{enumerate}
3022: 
3023: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3024: 
3025: \section*{Appendix B: More on Global Wavefunctions}
3026: 
3027: In this Appendix, we provide some technical details on the global wavefunctions
3028: $K_{\pm\pm,\vec{k}}$ used in Section~\ref{sec:string}.
3029: For simplicity of notation, we will suppress the subscript $\vec{k}$.
3030: As compared to \cite{Craps:2002ii,
3031: Elitzur:2002rt}, we restrict the discussion to the case of interest in this
3032: paper, as explained at the beginning of subsection~\ref{subsub:global}. In the
3033: notation of \cite{Craps:2002ii}, this means that
3034: \be\label{restr}
3035: \lambda=\mu=-j~,
3036: \ee
3037: with $j$ defined in \eq{eq:43} and \eq{eq:50} as
3038: \be\label{eq:50bis}
3039: j=-{1\over 2}+i{E_{\vec{k}}\over2Q}, \ \ \  E_{\vec k}=\sqrt{\vec k^2-Q^2}.
3040: \ee
3041: 
3042: We start by recalling some properties of the hypergeometric function
3043: $F(a,b;c;z)\equiv _2F_1(a,b;c;z)$ \cite{Abramowitz}. Using the notation
3044: \be\label{an}
3045: (a)_n={\G(a+n)\over\G(a)},
3046: \ee
3047: this function has the power series expansion
3048: \be\label{powerseries}
3049: F(a,b;c;z)=\sum_{n=0}^\infty{(a)_n(b)_n\over(c)_n}{z^n\over n!}
3050: %={\Gamma(c)
3051: %\over\Gamma(a)\Gamma(b)}\sum_{n=0}^\infty{\Gamma(a+n)\Gamma(b+n)\over\Gamma(c+n)}
3052: %{z^n\over n!}
3053: ,
3054: \ee
3055: which converges at least for $|z|<1$. It is useful to note that as $z\rightarrow 0$,
3056: \begin{equation}
3057: F(a,b;c;z)\longrightarrow 1
3058: \label{eq:AA}
3059: \end{equation}
3060: for all values of the parameters $a$, $b$ and $c$.
3061: The behavior for large $|z|$ can be obtained, for $b-a$ not integer, using the transformation formula
3062: \beqa\label{transfform}
3063: F(a,b;c;z)&=&{\G(c)\G(b-a)\over\G(b)\G(c-a)}(-z)^{-a}F(a,1-c+a;1-b+a;{1\over z})\cr
3064: &&+{\G(c)\G(a-b)\over\G(a)\G(c-b)}(-z)^{-b}F(b,1-c+b;1-a+b;{1\over z}), \ \ \
3065: \eeqa
3066: which is valid for $|\arg(-z)|<\pi$.
3067: Equation \eq{transfform} is singular for integer $b-a$. For $a=b$, it is replaced by
3068: \beqa\label{transfformaa}
3069: &&F(a,a;c;z)={\G(c)\over\G(a)\G(c-a)}(-z)^{-a}\sum_{n=0}^\infty
3070: {(a)_n(1-c+a)_n \over (n!)^2}z^{-n}\cr
3071: &&\ \ \ \ \ \ \ \ \ \ \ \ \  \times\left(\log(-z)+2\Psi(n+1)-\Psi(a+n)-\Psi(c-a-n)\right),\ \
3072: \eeqa
3073: where
3074: \be\label{Psi}
3075: \Psi(z)={\G'(z)\over\G(z)}.
3076: \ee
3077: Expression~\eq{transfformaa} is valid for $|\arg(-z)|<\pi$, $|z|>1$ and $c-a$ not integer.
3078: Other transformation formulas we will use are
3079: \be\label{transfformbis}
3080: F(a,b;c;z)=(1-z)^{-a}F\left(a,c-b;c;{z\over z-1}\right)
3081: \ee
3082: and
3083: \beqa\label{transfformtris}
3084: &&F(a,b;a+b;z)={\G(a+b)\over\G(a)\G(b)}\sum_{n=0}^\infty{(a)_n(b)_n\over(n!)^2}(1-z)^n\cr
3085: &&\ \ \ \ \ \ \ \ \ \ \ \ \times \left(-\log(1-z)+2\Psi(n+1)-\Psi(a+n)-\Psi(b+n)\right),\ \
3086: \eeqa
3087: the latter of which is valid for $|\arg(1-z)|<\pi$ and $|1-z|<1$. Also, recall that the Euler Beta function $B(a,b)$ is defined by
3088: \begin{equation}
3089: B(a,b)=\frac{\Gamma(a)\Gamma(b)}{\Gamma(a+b)}.
3090: \label{eq:new1}
3091: \end{equation}
3092: 
3093: 
3094: In terms of these functions, the global wavefunctions $K_{\pm\pm}$ are defined as follows.\\
3095: 
3096: \noindent Region I:
3097: \beqa\label{KI}
3098: K_{++}(uv)&=&{1\over2\pi i}B(-j,-j)(-uv)^j F\left(-j,-j;-2j;{1\over uv}\right)\cr
3099: K_{--}(uv)&=&{1\over2\pi i}B(j+1,j+1)(-uv)^{-j-1}F\left(j+1,j+1;2j+2;{1\over uv}\right)\cr
3100: K_{+-}(uv)&=&0\cr
3101: K_{-+}(uv)&=&{1\over\pi i}B(-j,j+1)F(-j,j+1;1;uv)
3102: \eeqa
3103: Region II:
3104: \beqa\label{KII}
3105: K_{++}(uv)&=&{1\over2\pi i}B(j+1,j+1)(-uv)^{-j-1}F\left(j+1,j+1;2j+2;{1\over uv}\right)\cr
3106: K_{--}(uv)&=&{1\over2\pi i}B(-j,-j)(-uv)^j F\left(-j,-j;-2j;{1\over uv}\right)\cr
3107: K_{+-}(uv)&=&{1\over\pi i}B(-j,j+1)F(-j,j+1;1;uv)\cr
3108: K_{-+}(uv)&=&0
3109: \eeqa
3110: Region III:
3111: \beqa\label{KIII}
3112: K_{++}(uv)&=&{1\over2\pi i}B(-j,j+1)(uv)^j F\left(-j,-j;1;1-{1\over uv}\right)\cr
3113: K_{--}(uv)&=&{1\over2\pi i}B(-j,j+1)(uv)^{-j-1}F\left(j+1,j+1;1;1-{1\over uv}\right)\cr
3114: K_{+-}(uv)&=&{1\over2\pi i}B(-j,j+1)(1-uv)^j F\left(-j,-j;1;{uv\over uv-1}\right)\cr
3115: K_{-+}(uv)&=&{1\over2\pi i}B(-j,j+1)(1-uv)^{-j-1}F\left(j+1,j+1;1;{uv\over uv-1}\right)\ \ \ \ \ \ \
3116: \eeqa
3117: Region IV:
3118: \beqa\label{KIV}
3119: K_{++}(uv)&=&{1\over2\pi i}B(-j,j+1)(uv)^{-j-1}F\left(j+1,j+1;1;1-{1\over uv}\right)\cr
3120: K_{--}(uv)&=&{1\over2\pi i}B(-j,j+1)(uv)^j F\left(-j,-j;1;1-{1\over uv}\right)\cr
3121: K_{+-}(uv)&=&{1\over2\pi i}B(-j,j+1)(1-uv)^{-j-1}F\left(j+1,j+1;1;{uv\over
3122: uv-1}\right)\cr
3123: K_{-+}(uv)&=&{1\over2\pi i}B(-j,j+1)(1-uv)^j F\left(-j,-j;1;1;{uv\over
3124: uv-1}\right)
3125: \eeqa
3126: Region V:
3127: \beqa\label{KV}
3128: K_{++}(uv)&=&{1\over\pi i}B(-j,j+1)F(-j,j+1;1;1-uv)\cr
3129: K_{--}(uv)&=&0\cr
3130: K_{+-}(uv)&=&{1\over2\pi i}B(j+1,j+1)(uv-1)^{-j-1}F\left(j+1,j+1;2j+2;{1\over 1-uv}\right)\cr
3131: K_{-+}(uv)&=&{1\over2\pi i}B(-j,-j)(uv-1)^j F\left(-j,-j;-2j;{1\over 1-uv}\right)
3132: \eeqa
3133: Region VI:
3134: \beqa\label{KVI}
3135: K_{++}(uv)&=&0\cr
3136: K_{--}(uv)&=&{1\over\pi i}B(-j,j+1)F(-j,j+1;1;1-uv)\cr
3137: K_{+-}(uv)&=&{1\over2\pi i}B(-j,-j)(uv-1)^j F\left(-j,-j;-2j;{1\over 1-uv}\right)\cr
3138: K_{-+}(uv)&=&{1\over2\pi i}B(j+1,j+1)(uv-1)^{-j-1}F\left(j+1,j+1;2j+2;
3139: {1\over 1-uv}\right)\cr &&
3140: \eeqa
3141: 
3142: First, we use these explicit formulas to determine the behavior of some of these
3143: wavefunctions in the asymptotic early and late time regimes. In
3144: regions I and II, we can use \eq{uvtx} to write
3145: \be\label{uvt}
3146: uv=-\sinh^2(Qt),
3147: \ee
3148: while in regions V and VI
3149: \be\label{uvtbis}
3150: 1-uv=-\sinh^2(Qt).
3151: \ee
3152: Let us concentrate on the early time part of Region I ($t\ll-1/Q$). It follows from
3153: \eq{KI}, \eq{uvt} and \eq{powerseries} that as $t\to-\infty$
3154: \be\label{K++early}
3155: K_{++}\rightarrow {1\over2\pi i}B(-j,-j)2^{-2j}e^{-2jQt}=
3156: {1\over2\pi i}B(-j,-j)4^{-j}e^{Qt}e^{-iE_{\vec{k}}t}.
3157: \ee
3158: This shows that $K_{++}$ reduces to a positive frequency wave for early times in
3159: Region
3160: I. Moreover, expression~\eq{K++early} allows us to determine the factor
3161:  $\N$ such that
3162: \be\label{calKnorm}
3163: \K_{++}=\N K_{++}
3164: \ee
3165: is normalized (see subsection~\ref{subsub:global}). The factor $\N$ is actually
3166: the same for all $K_{\pm\pm}$. Comparing \eq{K++early} with \eq{eq:48} or \eq{eq:49},
3167: we find
3168: \be\label{calN}
3169: \N={4^je^{\phi_0}2\pi i\over \sqrt {E_{\vec{k}}} B(-j,-j)}.
3170: \ee
3171: Using \eq{transfform}, we can rewrite $K_{-+}$ in Region I as
3172: \beqa\label{K-+I}
3173: K_{-+}&=&{1\over \pi i}B(-j,2j+1)(-uv)^jF\left(-j,-j;-2j;{1\over uv}\right)\cr
3174:       &&+{1\over \pi i}B(j+1,-2j-1)(-uv)^{-j-1}F\left(j+1,j+1;2j+2;{1\over uv}\right).\ \ \ \ \ \ \
3175: \eeqa
3176: This shows that asymptotically in Region~I, $K_{-+}$ is a superposition of positive
3177: and negative frequency waves with equal amplitudes.
3178: 
3179: Equivalently to \eq{K-+I}, we can rewrite $K_{+-}$ in Region II as
3180: \beqa\label{K+-II}
3181: K_{+-}&=&{1\over \pi i}B(-j,2j+1)(-uv)^jF\left(-j,-j;-2j;{1\over uv}\right)\cr
3182:       &&+{1\over \pi i}B(j+1,-2j-1)(-uv)^{-j-1}F\left(j+1,j+1;2j+2;{1\over uv}\right),\ \ \ \ \ \ \
3183: \eeqa
3184: which shows that asymptotically in Region~II, $K_{+-}$ is a superposition of positive
3185: and negative frequency waves with equal amplitudes.
3186: 
3187: Next, we discuss the behavior near the Big Crunch/Big Bang singularity $uv=0$. First
3188: consider $K_{++}$. In Region~I,
3189: we can use \eq{transfformaa} and \eq{KI} to write
3190: \be\label{K++Izero}
3191: K_{++}(uv)={1\over 2\pi i}\sum_{n=0}^\infty{(-j)_n(j+1)_n\over(n!)^2}(uv)^n(-\log(-uv)
3192: +2\Psi(n+1)-2\Psi(-j+n)).
3193: \ee
3194: We see that in Region I, $K_{++}$ diverges logarithmically near $uv=0$.
3195: Similarly, we find in Region~II
3196: \be\label{K++IIzero}
3197: K_{++}(uv)={1\over 2\pi i}\sum_{n=0}^\infty{(-j)_n(j+1)_n\over(n!)^2}(uv)^n(-\log(-uv)
3198: +2\Psi(n+1)-2\Psi(j+1+n)).
3199: \ee
3200: Applying \eq{transfformbis} and \eq{transfformtris} to \eq{KIII}, we find that
3201: in Region~III
3202: \beqa\label{K++IIIzero}
3203: &&K_{++}(uv)={1\over 2\pi i}\sum_{n=0}^\infty{(-j)_n(j+1)_n\over(n!)^2}(uv)^n\cr
3204: &&\ \ \ \ \ \ \ \ \ \ \ \  \ \times(-\log(uv)
3205: +2\Psi(n+1)-\Psi(-j+n)-\Psi(j+1+n)).\ \ \ \ \ \ \
3206: \eeqa
3207: The same equation, \eq{K++IIIzero}, holds in Region~IV as well.
3208: Now consider $K_{+-}$
3209: near $uv=0$. In Region~I, it vanishes. The small $uv$ behavior in Region~II is manifest
3210: from \eq{powerseries} and \eq{KII}. In Region~III, we can use \eq{transfformbis} to
3211: rewrite \eq{KIII} as
3212: \be\label{K+-IIIzero}
3213: K_{+-}(uv)={1\over2\pi i}B(-j,j+1)F(-j,j+1;1;uv).
3214: \ee
3215: The same formula, \eq{K+-IIIzero}, also holds in Region~IV.
3216: Therefore, as $uv$ approaches zero, $K_{+-}$ approaches a constant in regions
3217: III and IV, and twice that constant in
3218: Region~II.
3219: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3220: 
3221: \section*{Appendix C: Comparison with the Milne Orbifold}
3222: 
3223: In subsection~\ref{sub:string:fluctuations} we gave a qualitative discussion of some
3224: differences and similarities between the pure and generalized Milne orbifolds. This
3225: comparison can be made very concrete at the level of global wavefunctions, which we will do
3226: in this Appendix. Some of these observations were first made by the authors of
3227: \cite{Berkooz:2002je}. As in the previous Appendix, we will suppress the
3228: subscript $\vec{k}$ for notational simplicity.
3229: 
3230: In the limit that
3231: \be\label{limit}
3232: uv\to0 \ \ {\rm with} \quad {\vec k^2uv\over Q^2}\ \ {\rm and} \quad Q r_{0} \quad {\rm fixed},
3233: \ee
3234: where $r_{0}$ is defined in \eq{uvid}, the equation of motion \eq{eq:37}
3235: reduces to that of a massless scalar in the Milne orbifold \eq{eommilne}. We now
3236: take this limit of the solutions $\K_{++}$ and
3237: $\K_{+-}$ and see what they correspond to in the pure Milne case.
3238: First, note that in the limit \eq{limit},
3239: \be\label{limitEQ}
3240: {E_{\vec{k}}\over Q}\to+\infty.
3241: \ee
3242: It then follows from \eq{eq:50bis} that
3243: \be\label{limitj}
3244: j\to-{1\over 2}+i\infty.
3245: \ee
3246: In this limit, the normalization factor \eq{calN} becomes
3247: \be\label{Nlimit}
3248: |\N|\to{e^{\phi_0}\sqrt\pi\over\sqrt{8Q}}.
3249: \ee
3250: Using the fact that
3251: \be\label{Psiinfty}
3252: \Psi(z)\simeq\log(z)\ \ \ \ {\rm for}\ z\to\infty \ {\rm with}\
3253: |\arg(z)|<\pi,
3254: \ee
3255: we find from \eq{K++Izero} that in regions I and II
3256: \be\label{K++milne}
3257: K_{++}\to{1\over 2\pi
3258: i}\sum_{n=0}^\infty{1\over(n!)^2}\left({E_{\vec{k}}^2uv\over 4Q^2}\right)^n
3259: \left(-\log\left(-{E_{\vec{k}}^2uv\over4Q^2}\right)+\pi i+2\Psi(n+1)\right)
3260: \ee
3261: and
3262: \be\label{K++milneII}
3263: K_{++}\to{1\over 2\pi
3264: i}\sum_{n=0}^\infty{1\over(n!)^2}\left({E_{\vec{k}}^2uv\over 4Q^2}\right)^n
3265: \left(-\log\left(-{E_{\vec{k}}^2uv\over4Q^2}\right)-\pi i+2\Psi(n+1)\right)
3266: \ee
3267: respectively.
3268: Since in Region II $uv=-Q^2t^2$ in the limit \eq{limit},
3269: the right hand side of \eq{K++milneII} is precisely proportional to the
3270: Hankel function $H^{(2)}_0(Et)$, that is,
3271: \be\label{K++milneHankel}
3272: K_{++}\to-{1\over 2}H^{(2)}_0(E_{\vec{k}}t),
3273: \ee
3274: while \eq{K++milne} is its analytic continuation to Region I with
3275: negative $t$ via the lower half $t$ plane~\cite{Tolley:2002cv}.
3276: Therefore, the limit \eq{limit} of $\K_{++}$ in regions I and II is the
3277: wavefunction used in~\cite{Berkooz:2002je,  Tolley:2002cv} and~\cite{Nekrasov:2002kf} to define the adiabatic
3278: vacuum inherited from Minkowski space.
3279: In Region III, using \eq{K++IIIzero} and \eq{limitj}, we find that
3280: \be\label{K++milneIII}
3281: K_{++}\to{1\over 2\pi
3282: i}\sum_{n=0}^\infty{1\over(n!)^2}\left({E_{\vec{k}}^2uv\over 4Q^2}\right)^n
3283: \left(-\log\left({E_{\vec{k}}^2uv\over4Q^2}\right)+2\Psi(n+1)\right),
3284: \ee
3285: which is the expansion of the modified Bessel function $K_0$:
3286: \be\label{K++milnemodHankel}
3287: K_{++}\to{1\over \pi i}K_0\left(\sqrt{uvE_{\vec{k}}^2\over Q^2}\right).
3288: \ee
3289: The function $K_0$ decays asymptotically (that is, for large values of its argument).
3290: This is exactly the analytic continuation of \eq{K++milneHankel} to
3291: Region III and, additionally, to Region IV, via the lower half $t$ plane.
3292: We conclude that in all four regions of pure Milne space I, II, III and IV, the limit \eq{limit}
3293: of $\K_{++}$ exactly coincides with the wavefunctions used in
3294: \cite{Berkooz:2002je} to define the vacuum inherited from Minkowski space.
3295: These wavefunctions coincide with those of \cite{Tolley:2002cv} when
3296: restricted to regions I and II.
3297: 
3298: To determine the behavior of $\K_{+-}$ , note that in the limit~\eq{limit}
3299: \be\label{limithyper}
3300: F(-j,j+1;1;uv)\to\sum_{n=0}^\infty{1\over (n!)^2}\left({uvE_{\vec{k}}^2\over4Q^2}\right)^n.
3301: \ee
3302: It follows that in Region II
3303: \be\label{limithypertwo}
3304: F(-j,j+1;1;uv)\to J_0(E_{\vec{k}}t),
3305: \ee
3306: with $J_0$ a Bessel function,
3307: whereas in regions III and IV
3308: \be\label{limithyperthree}
3309: F(-j,j+1;1;uv)\to I_0\left(\sqrt{uvE_{\vec{k}}^2\over Q^2}\right),
3310: \ee
3311: with $I_0$ a modified Bessel function which has the property that it
3312: grows asymptotically. For this reason, $I_0$ it is usually not considered in
3313: discussions of the Milne orbifold, as we mentioned in
3314: subsection~\ref{sub:string:fluctuations}. In the generalized Milne orbifold, there is
3315: nothing wrong with growing behavior near $uv=0$ since regions III and IV do
3316: not extend to infinity. In the limit \eq{limit}, where regions III and IV are
3317: blown up to infinite size, the prefactor $B(-j,j+1)$ in \eq{KII} and \eq{K+-IIIzero}
3318: actually goes to zero. Therefore, although in regions III and IV $K_{+-}$ is proportional
3319: to \eq{limithyperthree} in the limit \eq{limit}, the proportionality factor is zero
3320: and $K_{+-}$ vanishes, as it does in regions I and II.
3321: 
3322: 
3323: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3324: \begin{thebibliography}{mm}
3325: 
3326: \bibitem{cosmoA}
3327: J.~Khoury, B.~A.~Ovrut, P.~J.~Steinhardt and N.~Turok,
3328: ``The ekpyrotic universe: Colliding branes and the origin of the hot big  bang,''
3329: Phys.\ Rev.\ D {\bf 64}, 123522 (2001)
3330: [hep-th/0103239].
3331: 
3332: \bibitem{cosmoB}
3333: R.~Y.~Donagi, J.~Khoury, B.~A.~Ovrut, P.~J.~Steinhardt and N.~Turok,
3334: ``Visible branes with negative tension in heterotic M-theory,''
3335: JHEP {\bf 0111}, 041 (2001)
3336: [hep-th/0105199].
3337: 
3338: \bibitem{cosmoC}
3339: J.~Khoury, B.~A.~Ovrut, P.~J.~Steinhardt and N.~Turok,
3340: ``Density perturbations in the ekpyrotic scenario,''
3341: Phys.\ Rev.\ D {\bf 66}, 046005 (2002)
3342: [hep-th/0109050].
3343: 
3344: %\cite{Gasperini:2002bn}
3345: \bibitem{Gasperini:2002bn}
3346: M.~Gasperini and G.~Veneziano,
3347: ``The pre-big bang scenario in string cosmology,''
3348: Phys.\ Rept.\  {\bf 373}, 1 (2003)
3349: [hep-th/0207130].
3350: %%CITATION = HEP-TH 0207130;%%
3351: 
3352: \bibitem{cosmoD}
3353: P.~Horava and E.~Witten,
3354: ``Heterotic and type I string dynamics from eleven dimensions,''
3355: Nucl.\ Phys.\ B {\bf 460}, 506 (1996)
3356: [hep-th/9510209];
3357: ``Eleven-Dimensional Supergravity on a Manifold with Boundary,''
3358: Nucl.\ Phys.\ B {\bf 475}, 94 (1996)
3359: [hep-th/9603142].
3360: 
3361: \bibitem{cosmoE}
3362: A.~Lukas, B.~A.~Ovrut and D.~Waldram,
3363: ``On the four-dimensional effective action of strongly coupled heterotic  string theory,''
3364: Nucl.\ Phys.\ B {\bf 532}, 43 (1998)
3365: [hep-th/9710208];
3366: ``The universe as a domain wall,''
3367: Phys.\ Rev.\ D {\bf 59}, 086001 (1999)
3368: [hep-th/9803235];
3369: ``Heterotic M-theory in five dimensions,''
3370: Nucl.\ Phys.\ B {\bf 552}, 246 (1999)
3371: [hep-th/9806051].
3372: 
3373: %\cite{Khoury:2001bz}
3374: \bibitem{Khoury:2001bz}
3375: J.~Khoury, B.~A.~Ovrut, N.~Seiberg, P.~J.~Steinhardt and N.~Turok,
3376: ``From Big Crunch to Big Bang,''
3377: Phys.\ Rev.\ D {\bf 65}, 086007 (2002)
3378: [hep-th/0108187].
3379: 
3380: \bibitem{cosmoF}
3381: P.~J.~Steinhardt and N.~Turok,
3382: ``Cosmic evolution in a cyclic universe,''
3383: Phys.\ Rev.\ D {\bf 65}, 126003 (2002)
3384: [hep-th/0111098];
3385: ``Is vacuum decay significant in ekpyrotic and cyclic models?,''
3386: Phys.\ Rev.\ D {\bf 66}, 101302 (2002)
3387: [astro-ph/0112537];
3388: S.~Gratton, J.~Khoury, P.~J.~Steinhardt and N.~Turok,
3389: ``Conditions for generating scale-invariant density perturbations,''
3390: astro-ph/0301395;
3391: J.~Khoury, P.~J.~Steinhardt and N.~Turok,
3392: ``Great expectations: Inflation versus cyclic predictions for spectral  tilt,''
3393: astro-ph/0302012;
3394: ``Designing cyclic universe models,''
3395: hep-th/0307132.
3396: 
3397: \bibitem{no}
3398: D.~H.~Lyth,
3399: ``The primordial curvature perturbation in the ekpyrotic universe,''
3400: Phys.\ Lett.\ B {\bf 524}, 1 (2002)
3401: [hep-ph/0106153];
3402: R.~Brandenberger and F.~Finelli,
3403: ``On the spectrum of fluctuations in an effective field theory of the  ekpyrotic universe,''
3404: JHEP {\bf 0111}, 056 (2001)
3405: [hep-th/0109004];
3406: ``On the generation of a scale-invariant spectrum of adiabatic  fluctuations in cosmological models with a contracting phase,''
3407: Phys.\ Rev.\ D {\bf 65}, 103522 (2002)
3408: [hep-th/0112249];
3409: J.~c.~Hwang,
3410: ``Cosmological structure problem in the ekpyrotic scenario,''
3411: Phys.\ Rev.\ D {\bf 65}, 063514 (2002)
3412: [astro-ph/0109045].
3413: 
3414: \bibitem{maybe}
3415: C.~Cartier, R.~Durrer and E.~J.~Copeland,
3416: ``Cosmological perturbations and the transition from contraction to  expansion,''
3417: Phys.\ Rev.\ D {\bf 67}, 103517 (2003)
3418: [hep-th/0301198].
3419: 
3420: \bibitem{Tolley:2002cv}
3421: A.~J.~Tolley and N.~Turok,
3422: ``Quantum fields in a big crunch / big bang spacetime,''
3423: Phys.\ Rev.\ D {\bf 66}, 106005 (2002)
3424: [hep-th/0204091].
3425: 
3426: \bibitem{Tolley:2003nx}
3427: A.~J.~Tolley, N.~Turok and P.~J.~Steinhardt,
3428: ``Cosmological perturbations in a big crunch / big bang space-time,''
3429: hep-th/0306109.
3430: %%CITATION = HEP-TH 0306109;%%
3431: 
3432: %\cite{Myers:fv}
3433: \bibitem{Myers:fv}
3434: R.~C.~Myers,
3435: ``New Dimensions For Old Strings,''
3436: Phys.\ Lett.\ B {\bf 199}, 371 (1987).
3437: %%CITATION = PHLTA,B199,371;%%
3438: 
3439: %\cite{Antoniadis:1988vi}
3440: \bibitem{Antoniadis:1988vi}
3441: I.~Antoniadis, C.~Bachas, J.~R.~Ellis and D.~V.~Nanopoulos,
3442: ``An Expanding Universe In String Theory,''
3443: Nucl.\ Phys.\ B {\bf 328}, 117 (1989).
3444: %%CITATION = NUPHA,B328,117;%%
3445: 
3446: %\cite{Craps:2002ii}
3447: \bibitem{Craps:2002ii}
3448: B.~Craps, D.~Kutasov and G.~Rajesh,
3449: ``String propagation in the presence of cosmological singularities,''
3450: JHEP {\bf 0206}, 053 (2002)
3451: [hep-th/0205101].
3452: %%CITATION = HEP-TH 0205101;%%
3453: 
3454: %\cite{Witten:1991yr}
3455: \bibitem{Witten:1991yr}
3456: E.~Witten,
3457: ``On string theory and black holes,''
3458: Phys.\ Rev.\ D {\bf 44}, 314 (1991).
3459: %%CITATION = PHRVA,D44,314;%%
3460: 
3461: %\cite{Maldacena:2001kr}
3462: \bibitem{Maldacena:2001kr}
3463: J.~M.~Maldacena,
3464: ``Eternal black holes in Anti-de-Sitter,''
3465: JHEP {\bf 0304}, 021 (2003)
3466: [hep-th/0106112].
3467: %%CITATION = HEP-TH 0106112;%%
3468: 
3469: \bibitem{BTZ}
3470: S.~Hemming, E.~Keski-Vakkuri and P.~Kraus,
3471: ``Strings in the extended BTZ spacetime,''
3472: JHEP {\bf 0210}, 006 (2002)
3473: [hep-th/0208003];
3474: P.~Kraus, H.~Ooguri and S.~Shenker,
3475: ``Inside the horizon with AdS/CFT,''
3476: Phys.\ Rev.\ D {\bf 67}, 124022 (2003)
3477: [hep-th/0212277];
3478: T.~S.~Levi and S.~F.~Ross,
3479: ``Holography beyond the horizon and cosmic censorship,''
3480: hep-th/0304150;
3481: L.~Fidkowski, V.~Hubeny, M.~Kleban and S.~Shenker,
3482: ``The black hole singularity in AdS/CFT,''
3483: hep-th/0306170.
3484: 
3485: %\cite{Seiberg:2002hr}
3486: \bibitem{Seiberg:2002hr}
3487: N.~Seiberg,
3488: ``From big crunch to big bang - is it possible?,''
3489: hep-th/0201039.
3490: %%CITATION = HEP-TH 0201039;%%
3491: 
3492: 
3493: %\cite{Elitzur:2002rt}
3494: \bibitem{Elitzur:2002rt}
3495: S.~Elitzur, A.~Giveon, D.~Kutasov and E.~Rabinovici,
3496: ``From big bang to big crunch and beyond,''
3497: JHEP {\bf 0206}, 017 (2002)
3498: [hep-th/0204189].
3499: %%CITATION = HEP-TH 0204189;%%
3500: 
3501: %\cite{Berkooz:2002je}
3502: \bibitem{Berkooz:2002je}
3503: M.~Berkooz, B.~Craps, D.~Kutasov and G.~Rajesh,
3504: ``Comments on cosmological singularities in string theory,''
3505: JHEP {\bf 0303}, 031 (2003)
3506: [hep-th/0212215].
3507: %%CITATION = HEP-TH 0212215;%%
3508: 
3509: %\cite{Liu:2002ft}
3510: \bibitem{Liu:2002ft}
3511: H.~Liu, G.~Moore and N.~Seiberg,
3512: ``Strings in a time-dependent orbifold,''
3513: JHEP {\bf 0206}, 045 (2002)
3514: [hep-th/0204168].
3515: %%CITATION = HEP-TH 0204168;%%
3516: 
3517: %\cite{Horowitz:2002mw}
3518: \bibitem{Horowitz:2002mw}
3519: G.~T.~Horowitz and J.~Polchinski,
3520: ``Instability of spacelike and null orbifold singularities,''
3521: Phys.\ Rev.\ D {\bf 66}, 103512 (2002)
3522: [hep-th/0206228].
3523: %%CITATION = HEP-TH 0206228;%%
3524: 
3525: \bibitem{Cornalba:2002fi}
3526: L.~Cornalba and M.~S.~Costa,
3527: ``A New Cosmological Scenario in String Theory,''
3528: Phys.\ Rev.\ D {\bf 66}, 066001 (2002)
3529: [hep-th/0203031].
3530: 
3531: \bibitem{related}
3532: J.~Simon,
3533: ``The geometry of null rotation identifications,''
3534: JHEP {\bf 0206}, 001 (2002)
3535: [hep-th/0203201];
3536: L.~Cornalba, M.~S.~Costa and C.~Kounnas,
3537: ``A resolution of the cosmological singularity with orientifolds,''
3538: Nucl.\ Phys.\ B {\bf 637}, 378 (2002)
3539: [hep-th/0204261].
3540: 
3541: 
3542: \bibitem{backreaction}
3543: G.~T.~Horowitz and A.~R.~Steif,
3544: ``Singular String Solutions With Nonsingular Initial Data,''
3545: Phys.\ Lett.\ B {\bf 258}, 91 (1991);
3546: S.~W.~Hawking,
3547: ``The Chronology protection conjecture,''
3548: Phys.\ Rev.\ D {\bf 46}, 603 (1992);
3549: A.~Lawrence,
3550: ``On the instability of 3D null singularities,''
3551: JHEP {\bf 0211}, 019 (2002)
3552: [hep-th/0205288];
3553: E.~J.~Martinec and W.~McElgin,
3554: ``Exciting AdS orbifolds,''
3555: JHEP {\bf 0210}, 050 (2002)
3556: [hep-th/0206175];
3557: H.~Liu, G.~Moore and N.~Seiberg,
3558: ``Strings in time-dependent orbifolds,''
3559: JHEP {\bf 0210}, 031 (2002)
3560: [hep-th/0206182];
3561: M.~Fabinger and J.~McGreevy,
3562: ``On smooth time-dependent orbifolds and null singularities,''
3563: JHEP {\bf 0306}, 042 (2003)
3564: [hep-th/0206196];
3565: M.~Fabinger and S.~Hellerman,
3566: ``Stringy resolutions of null singularities,''
3567: hep-th/0212223;
3568: S.~Elitzur, A.~Giveon and E.~Rabinovici,
3569: ``Removing singularities,''
3570: JHEP {\bf 0301}, 017 (2003)
3571: [hep-th/0212242];
3572: L.~Cornalba and M.~S.~Costa,
3573: ``On the classical stability of orientifold cosmologies,''
3574: hep-th/0302137;
3575: A.~Giveon, E.~Rabinovici and A.~Sever,
3576: ``Strings in singular time-dependent backgrounds,''
3577: Fortsch.\ Phys.\  {\bf 51}, 805 (2003)
3578: [hep-th/0305137].
3579: 
3580: %\cite{Liu:2002yd}
3581: \bibitem{Liu:2002yd}
3582: H.~Liu, G.~Moore and N.~Seiberg,
3583: ``The challenging cosmic singularity,''
3584: gr-qc/0301001.
3585: %%CITATION = GR-QC 0301001;%%
3586: 
3587: 
3588: %\cite{Bachas:2002qt}
3589: \bibitem{Bachas:2002qt}
3590: C.~Bachas and C.~Hull,
3591: ``Null brane intersections,''
3592: JHEP {\bf 0212}, 035 (2002)
3593: [hep-th/0210269].
3594: %%CITATION = HEP-TH 0210269;%%
3595: 
3596: %\cite{Berkooz:2003bs}
3597: \bibitem{Berkooz:2003bs}
3598: M.~Berkooz and B.~Pioline,
3599: ``Strings in an electric field, and the Milne universe,''
3600: hep-th/0307280.
3601: %%CITATION = HEP-TH 0307280;%%
3602: 
3603: 
3604: %\cite{Nekrasov:2002kf}
3605: \bibitem{Nekrasov:2002kf}
3606: N.~A.~Nekrasov,
3607: ``Milne universe, tachyons, and quantum group,''
3608: hep-th/0203112.
3609: %%CITATION = HEP-TH 0203112;%%
3610: 
3611: 
3612: \bibitem{recent}
3613: V.~Balasubramanian, S.~F.~Hassan, E.~Keski-Vakkuri and A.~Naqvi,
3614: ``A space-time orbifold: A toy model for a cosmological singularity,''
3615: Phys.\ Rev.\ D {\bf 67}, 026003 (2003)
3616: [hep-th/0202187];
3617: S.~Kachru and L.~McAllister,
3618: ``Bouncing brane cosmologies from warped string compactifications,''
3619: JHEP {\bf 0303}, 018 (2003)
3620: [hep-th/0205209];
3621: A.~Buchel, P.~Langfelder and J.~Walcher,
3622: ``On time-dependent backgrounds in supergravity and string theory,''
3623: Phys.\ Rev.\ D {\bf 67}, 024011 (2003)
3624: [hep-th/0207214];
3625: A.~Hashimoto and S.~Sethi,
3626: ``Holography and string dynamics in time-dependent backgrounds,''
3627: Phys.\ Rev.\ Lett.\  {\bf 89}, 261601 (2002)
3628: [hep-th/0208126];
3629: J.~Simon,
3630: ``Null orbifolds in AdS, time dependence and holography,''
3631: JHEP {\bf 0210}, 036 (2002)
3632: [hep-th/0208165];
3633: M.~Alishahiha and S.~Parvizi,
3634: ``Branes in time-dependent backgrounds and AdS/CFT correspondence,''
3635: JHEP {\bf 0210}, 047 (2002)
3636: [hep-th/0208187];
3637: Y.~Satoh and J.~Troost,
3638: ``Massless BTZ black holes in minisuperspace,''
3639: JHEP {\bf 0211}, 042 (2002)
3640: [hep-th/0209195].
3641: R.~G.~Cai, J.~X.~Lu and N.~Ohta,
3642: ``NCOS and D-branes in time-dependent backgrounds,''
3643: Phys.\ Lett.\ B {\bf 551}, 178 (2003)
3644: [hep-th/0210206];
3645: K.~Okuyama,
3646: ``D-branes on the null-brane,''
3647: JHEP {\bf 0302}, 043 (2003)
3648: [hep-th/0211218];
3649: G.~Papadopoulos, J.~G.~Russo and A.~A.~Tseytlin,
3650: ``Solvable model of strings in a time-dependent plane-wave background,''
3651: Class.\ Quant.\ Grav.\  {\bf 20}, 969 (2003)
3652: [hep-th/0211289];
3653: B.~Durin and B.~Pioline,
3654: ``Open strings in relativistic ion traps,''
3655: JHEP {\bf 0305}, 035 (2003)
3656: [hep-th/0302159];
3657: J.~R.~David,
3658: ``Plane waves with weak singularities,''
3659: hep-th/0303013;
3660: R.~Biswas, E.~Keski-Vakkuri, R.~G.~Leigh, S.~Nowling and E.~Sharpe,
3661: ``The taming of closed time-like curves,''
3662: hep-th/0304241;
3663: J.~G.~Russo,
3664: ``Cosmological string models from Milne spaces and SL(2,Z) orbifold,''
3665: hep-th/0305032.
3666: 
3667: \bibitem{TseytlinJackJonesPanvel}
3668: A.~A.~Tseytlin,
3669: ``On the form of the black hole solution in D = 2 theory,''
3670: Phys.\ Lett.\ B {\bf 268}, 175 (1991);
3671: I.~Jack, D.~R.~Jones and J.~Panvel,
3672: ``Exact bosonic and supersymmetric string black hole solutions,''
3673: Nucl.\ Phys.\ B {\bf 393}, 95 (1993)
3674: [hep-th/9201039].
3675: 
3676: \bibitem{coset}
3677: A.~A.~Tseytlin and C.~Vafa,
3678: ``Elements of string cosmology,''
3679: Nucl.\ Phys.\ B {\bf 372}, 443 (1992)
3680: [hep-th/9109048];
3681: I.~Bars and K.~Sfetsos,
3682: ``Global analysis of new gravitational singularities in string and particle theories,''
3683: Phys.\ Rev.\ D {\bf 46}, 4495 (1992)
3684: [hep-th/9205037];
3685: ``Conformally exact metric and dilaton in string theory on curved space-time,''
3686: Phys.\ Rev.\ D {\bf 46}, 4510 (1992)
3687: [hep-th/9206006];
3688: D.~Lust,
3689: ``Cosmological string backgrounds,''
3690: hep-th/9303175.
3691: 
3692: %\cite{Kounnas:1992wc}
3693: \bibitem{Kounnas:1992wc}
3694: C.~Kounnas and D.~Lust,
3695: ``Cosmological string backgrounds from gauged WZW models,''
3696: Phys.\ Lett.\ B {\bf 289}, 56 (1992)
3697: [hep-th/9205046].
3698: %%CITATION = HEP-TH 9205046;%%
3699: 
3700: %\cite{Dijkgraaf:1991ba}
3701: \bibitem{Dijkgraaf:1991ba}
3702: R.~Dijkgraaf, H.~Verlinde and E.~Verlinde,
3703: ``String propagation in a black hole geometry,''
3704: Nucl.\ Phys.\ B {\bf 371}, 269 (1992).
3705: %%CITATION = NUPHA,B371,269;%%
3706: 
3707: \bibitem{Vilenkin}
3708: N.~J.~Vilenkin,
3709: ``Special Functions and the Theory of Group Representations,''
3710: AMS, 1968.
3711: 
3712: \bibitem{orb}
3713: L.~J.~Dixon, J.~A.~Harvey, C.~Vafa and E.~Witten,
3714: ``Strings On Orbifolds,''
3715: Nucl.\ Phys.\ B {\bf 261}, 678 (1985);
3716: ``Strings On Orbifolds. 2,''
3717: Nucl.\ Phys.\ B {\bf 274}, 285 (1986);
3718: L.~J.~Dixon, D.~Friedan, E.~J.~Martinec and S.~H.~Shenker,
3719: ``The Conformal Field Theory Of Orbifolds,''
3720: Nucl.\ Phys.\ B {\bf 282}, 13 (1987).
3721: 
3722: \bibitem{grav}
3723: L.~H.~Ford,
3724: ``Gravitational Particle Creation And Inflation,''
3725: Phys.\ Rev.\ D {\bf 35}, 2955 (1987);
3726: B.~Spokoiny,
3727: ``Deflationary universe scenario,''
3728: Phys.\ Lett.\ B {\bf 315}, 40 (1993)
3729: [gr-qc/9306008];
3730: P.~J.~Peebles and A.~Vilenkin,
3731: ``Quintessential inflation,''
3732: Phys.\ Rev.\ D {\bf 59}, 063505 (1999)
3733: [astro-ph/9810509].
3734: 
3735: \bibitem{Abramowitz}
3736: M.~Abramowitz and I.~A.~Stegun,
3737: ``Handbook of Mathematical Functions,''
3738: Dover Publications.
3739: 
3740: 
3741: 
3742: \end{thebibliography}
3743: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3744: 
3745: 
3746: 
3747: 
3748: 
3749: 
3750: 
3751: 
3752: 
3753: 
3754: 
3755: 
3756: 
3757: 
3758: 
3759: 
3760: 
3761: \end{document}
3762: