hep-th0302137/HP.tex
1: \documentclass[11pt]{article}
2: \usepackage{amssymb}
3: 
4: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5: \usepackage{graphicx}
6: \usepackage{amsmath}
7: 
8: %TCIDATA{OutputFilter=LATEX.DLL}
9: %TCIDATA{Created=Thu Jan 16 16:58:19 2003}
10: %TCIDATA{LastRevised=Wed Jan 22 14:41:53 2003}
11: %TCIDATA{<META NAME="GraphicsSave" CONTENT="32">}
12: %TCIDATA{<META NAME="DocumentShell" CONTENT="General\Blank Document">}
13: %TCIDATA{Language=American English}
14: %TCIDATA{CSTFile=LaTeX article (bright).cst}
15: 
16: \newtheorem{theorem}{Theorem}
17: \newtheorem{acknowledgement}[theorem]{Acknowledgement}
18: \newtheorem{algorithm}[theorem]{Algorithm}
19: \newtheorem{axiom}[theorem]{Axiom}
20: \newtheorem{case}[theorem]{Case}
21: \newtheorem{claim}[theorem]{Claim}
22: \newtheorem{conclusion}[theorem]{Conclusion}
23: \newtheorem{condition}[theorem]{Condition}
24: \newtheorem{conjecture}[theorem]{Conjecture}
25: \newtheorem{corollary}[theorem]{Corollary}
26: \newtheorem{criterion}[theorem]{Criterion}
27: \newtheorem{definition}[theorem]{Definition}
28: \newtheorem{example}[theorem]{Example}
29: \newtheorem{exercise}[theorem]{Exercise}
30: \newtheorem{lemma}[theorem]{Lemma}
31: \newtheorem{notation}[theorem]{Notation}
32: \newtheorem{problem}[theorem]{Problem}
33: \newtheorem{proposition}[theorem]{Proposition}
34: \newtheorem{remark}[theorem]{Remark}
35: \newtheorem{solution}[theorem]{Solution}
36: \newtheorem{summary}[theorem]{Summary}
37: \newenvironment{proof}[1][Proof]{\textbf{#1.} }{\ \rule{0.5em}{0.5em}}
38: \input{tcilatex}
39: 
40: \begin{document}
41: 
42: \vspace*{-.6in} \thispagestyle{empty}
43: \begin{flushright}
44: ITFA--2003--05\\
45: CFP--2003--01\\
46: \end{flushright}
47: \vspace{.2in} {\Large
48: \begin{center}
49: {\bf On the Classical Stability of Orientifold Cosmologies}
50: \end{center}}
51: \vspace{.2in}
52: \begin{center}
53: \textbf{Lorenzo Cornalba}$^{\dagger}$\footnote{lcornalb@science.uva.nl}\ \ 
54: and\ \ \textbf{Miguel S. Costa}$^{\ddagger}$\footnote{miguelc@fc.up.pt}
55: \\
56: \vspace{.2in} $^{\dagger}$\emph{Instituut voor Theoretische Fysica,
57: Universiteit van Amsterdam\\Valckenierstraat 65, 1018 XE Amsterdam, The Netherlands}\\
58: \vspace{.1in}
59: $^{\ddagger}$\emph{Centro de  F\'\i sica do Porto, Departamento de F\'\i sica\\ 
60: Faculdade de Ci\^encias, Universidade do Porto\\
61: Rua do Campo Alegre, 687 - 4169-007 Porto, Portugal
62: }
63: \end{center}
64: 
65: \vspace{.3in}
66: 
67: \begin{abstract}
68: 
69: \textit{We analyze the classical stability of string cosmologies driven by the dynamics of
70: orientifold planes. These models are related to time--dependent orbifolds, and resolve
71: the orbifold singularities which are otherwise problematic by introducing orientifold planes.
72: In particular, we show that the instability discussed by Horowitz and Polchinski
73: for pure orbifold models is resolved by the presence of
74: the orientifolds. Moreover, we discuss the issue of stability of the
75: cosmological Cauchy horizon, and we show that it is stable to small perturbations
76: due to in--falling matter.}
77: 
78: \end{abstract}
79: \newpage
80: 
81: \section{Introduction}
82: 
83: Over the past year there has been a renewed interest in the study of time--dependent string
84: backgrounds [1--33]. The main objective of this programme is to learn about the origin of the
85: (possibly apparent) cosmological singularity using the powerful techniques of string theory. The logic
86: behind this investigation was to start with an exact conformal field theory, \textit{e.g.} strings in
87: flat space, and to consider an orbifold of the theory with a time--dependent quotient space \cite{HS}. 
88: These simple orbifold constructions were, however, quite generally shown to be unstable by
89: Horowitz and Polchinski \cite{HP} 
90: \footnote{The exception is the null--brane \cite{SimonFigueroa} with extra non--compact directions \cite{LMS2}, 
91: which is an exact and regular bounce in string theory free of singularities. Notice, however, 
92: that this space is not a cosmology since it is supersymmetric and depends on a null, 
93: not a time--like, direction.}. These instabilities are 
94: associated to the formation of large black holes whenever a particle is coupled to the geometry
95: and are non-perturbative in the string coupling. For this reason a perturbative string resolution of 
96: the time--dependent orbifold singularities seems inappropriate. 
97: 
98: In \cite{CC} we investigated the orbifold of flat space--time by a boost and translation
99: transformation. This orbifold suffers from the Horowitz--Polchinski instability. However, 
100: an aditional essential  ingredient of the model was introduced in \cite{CCK}:
101: \textit{the time--like orbifold singularity is an orientifold, i.e. a brane with negative tension}.
102: One of the results that we shall show is that the orientifold provides precisely the non--perturbative
103: string resolution of the singularity, modifying the gravitational interaction in its proximity.
104: In fact, after briefly revising the Horowitz--Polchinski argument and the orientifold cosmology 
105: construction, we shall show in Section 4, with a tractable example in dimension $3$,
106: that \textit{due to the presence of the orientifolds large black holes are not formed.}
107: 
108: In any pre big--bang scenario, there is generically a possible classic instability associated
109: with the propagation of matter through the bounce, which is analogous to the propagation of matter 
110: between the outer and inner horizons of charged black--holes \cite{PenroseSimpson,CH}.
111: Let us consider a perturbation at some finite time in the far past, when the universe is contracting.
112: In the case of orientifold cosmology 
113: there is a future cosmological horizon where the contracting phase ends 
114: (see also \cite{KounnasLust,Q0,Q1}), therefore one needs to worry about such perturbations 
115: diverging at the horizon, thus creating a big crunch space--time.
116: We shall show in Section 5 that such perturbations remain finite (of course, they can grow and 
117: create small black holes, but this does not change the causal cosmological structure of the geometry). 
118: Furthermore, we shall see that the fluctuations that destabilize the geometry are precisely those 
119: that already destabilize the Minkowski vacuum.
120: 
121: \section{The Horowitz--Polchinski Problem\label{HPproblem}}
122: 
123: In \cite{HP}, Horowitz and Polchinski argue that a large class of time
124: dependent orbifolds are unstable to small perturbations, due to large
125: backreactions of the geometry. Their results, which overlap with the string
126: theory scattering computations of \cite{LMS1,LMS2}, do not rely on string
127: theory arguments, and are obtained simply within the framework of classical
128: general relativity.
129: 
130: The argument of Horowitz and Polchinski is quite simple and runs as follows.
131: They analyze the orbifold of $D$--dimensional Minkowski space $\mathbb{M}^{D}
132: $ by the action of a discrete isometry $e^{\kappa }$ where $\kappa $ is a
133: Killing vector, and consider, in the quotient orbifold space $\mathbb{M}%
134: ^{D}/e^{\kappa }$, a small perturbation. This perturbation naturally
135: corresponds to an infinite sequence of perturbations in the covering space $%
136: \mathbb{M}^{D}$, related to each other by the action of $e^{\kappa }$. They
137: consider, quite generally, a light ray with world--line $\Omega \subset 
138: \mathbb{M}^{3}$, together with all of its images 
139: \begin{equation*}
140: \Omega _{n}=\,e^{n\kappa }\Omega \,\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
141: \ \ \ \ \left( n\in \mathbb{Z}\right) 
142: \end{equation*}
143: for any integer $n$. The basic claim of \cite{HP} is that, for a wide range
144: of time dependent orbifolds, the backreaction of gravity to the presence
145: of this infinite sequence of light rays $\Omega _{n}$ will inevitably
146: produce large black holes. As we already mentioned, this statement is a
147: claim in pure classical gravity theory and to prove it Horowitz and
148: Polchinski consider the scattering of two light rays $\Omega _{0}$ and $%
149: \Omega _{n}$. They compute, given $\kappa $, the impact parameter $b_{n}$
150: and the center of mass energy $\mathcal{E}_{n}$ of the scattering, and show
151: that, for generic time dependent orbifolds, one has that $G\mathcal{E}%
152: _{n}\gg b_{n}^{D-3}$ for $n$ sufficiently large. The two image particles
153: will be well within the Schwarzchild radius corresponding to the center of
154: mass energy, and will form a black hole.
155: 
156: It is quite clear that, if the above argument is correct, it must hold also
157: for a continuous distribution of light rays 
158: \begin{equation}
159: e^{t\kappa }\Omega \,\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
160: \left( t\in \mathbb{R}\right)   \label{HPsurface}
161: \end{equation}
162: since the problematic interactions occur between rays separated by $\Delta
163: t\gg 1$, and therefore should be completely insensitive to the distribution
164: of the essentially parallel light rays when the parameter $t$ varies by $%
165: \Delta t\sim 1$. For continuous values of $t$, equation ($\ref{HPsurface}$)
166: defines a surface $S$ in $\mathbb{M}^{D}$, which we call the
167: Horowitz--Polchinski (HP) surface. Again, the backreaction of gravity to the
168: matter distribution on $S$ should form black holes by the previous argument.
169: On the other hand, the continuous distribution is easier to treat
170: analytically since it is invariant under the action of $\kappa $. Therefore,
171: the full solution with the gravitational backreaction will also have $\kappa 
172: $ as a Killing vector. This fact simplifies explicit computations and
173: therefore, from now on, we will only consider the continuous case.
174: 
175: In the rest of the paper we will be interested exclusively in a specific
176: choice of orbifold, which describes, when correctly interpreted in $M$%
177: --theory, the dynamics of an orientifold/anti--orientifold pair. We postpone
178: the discussion of the correct $M$--theory embedding and the crucial distinction
179: between the orbifold and the orientifold interpretation to section 
180: \ref{orientifold}, and we describe now the orbifold along the lines of \cite
181: {HP}.
182: 
183: We write $\mathbb{M}^{D}=\mathbb{M}^{3}\times \mathbb{E}^{D-3}$ and we
184: parametrize $\mathbb{M}^{3}$ with coordinates $X^{\pm }$, $Y$, with line element
185: \begin{equation}
186: -dX^{+}dX^{-}+dY^{2}\,.  \label{flat}
187: \end{equation}
188: Points in $\mathbb{M}^{D}$ are identified along the orbits of the Killing
189: vector field $\kappa $ which corresponds to boost in the $X^{\pm }$ plane
190: and translation in the orthogonal direction $Y$ 
191: \begin{equation}
192: \kappa =X^{+}\partial _{+}-X^{-}\partial _{-}+\,\partial _{Y}\,.
193: \label{killing}
194: \end{equation}
195: For this special choice of $\kappa $ one has $b_{n}\sim n$ and $\mathcal{E}%
196: _{n}\sim e^{n}$ and the bound $G\mathcal{E}_{n}\gg b_{n}^{D-3}$ is always
197: satisfied for $n$ large \cite{HP}. The directions along $\mathbb{E}^{D-3}$
198: then play no role (except to determine the strength of the gravitational
199: interaction) and we will omit any reference to them, working effectively in
200: three dimensions.
201: 
202: To study the geometry of the HP surface $S$, we parametrize the light ray $%
203: \Omega $ as 
204: \begin{equation*}
205: A^{a}+sV^{a}\text{\thinspace },
206: \end{equation*}
207: where $V^{a}$ is a null vector and $s$ is an affine parameter. If we
208: consider the generic case\footnote{%
209: When $V^{+}V^{-}=0$ the images $e^{t\kappa }\Omega $ correspond to parallel
210: light rays, and do not form black holes.} when $V^{+}V^{-}\neq 0$ then, for
211: some value of $t$, the image ray $e^{t\kappa }\Omega $ will be directed
212: along the null vector\footnote{%
213: It could also be directed along $\left( 1,1,-1\right) $. This case is just
214: the mirror under $X^{\pm }\rightarrow X^{\mp }$ and we do not discuss it.} $%
215: \left( 1,1,1\right) $. This specific light ray on the surface $S$ will
216: intersect the plane $X^{+}+X^{-}=0$ at a point with $X^{+}-X^{-}=2a$. By
217: shifting the variables $s,t,Y$ so that this point corresponds to $s=t=Y=0$
218: we see that the surface $S$ is given by the following parametrization 
219: \begin{eqnarray}
220: X^{\pm } &=&\left( s\pm a\right) e^{\pm t}\, ,  \label{surface} \\
221: Y &=&\left( s+t\right)\, .  \notag
222: \end{eqnarray}
223: \begin{figure}
224: \begin{center}
225: \includegraphics{fig7.eps}
226: \end{center}
227: \caption{The curves given by equation (\ref{surxpxm}),
228: which represent the surfaces $S_a$ for the values of $a$ 
229: indicated in the figure. The surface $S_1$ is clearly
230: singled out, and is the unique null surface.\label{fig7}}
231: \end{figure}
232: The only free parameter which labels different surfaces is the constant $a$,
233: and we will add a label $S_{a}$ to distinguish them. A convenient way to
234: visualize $S_{a}$ is to move to coordinates 
235: \begin{eqnarray}
236: x^{\pm } &=&X^{\pm }e^{\mp Y}\, ,  \label{2dcoord} \\
237: y &=&Y\, ,  \notag
238: \end{eqnarray}
239: so that $\kappa =\frac{\partial }{\partial y}$. Then, since the surface $%
240: S_{a}$ is invariant under $\kappa $, it is given by a curve in the $%
241: x^{+},x^{-}$ plane (see Figure \ref{fig7})
242: \begin{equation}
243: x^{\pm }=\left( s\pm a\right) e^{\mp s}.  \label{surxpxm}
244: \end{equation}
245: This curve can also be viewed as the intersection of the surface $S_{a}$
246: with the plane $Y=0$ in $\mathbb{M}^{3}$. 
247: 
248: 
249: Let us now describe some basic properties of the surfaces $S_{a}$. The
250: induced metric is given by 
251: \begin{equation*}
252: 2\left( 1-a\right) dsdt+\left( s^{2}-a^{2}+1\right) dt^{2}
253: \end{equation*}
254: so it is timelike for all values of $a$ except for $a=1$, when $S$ is a
255: null surface. For $a\neq 1$ the surfaces are smooth for any value of the
256: parameters $s,t$. For $a=1$ the surface is, on the other hand, divided into
257: two smooth parts for $s>0$ and $s<0$ which are joined along the singular
258: line $s=0$. Moreover, the forward directed light cones, which are tangent to
259: the null surface $S_{1}$, lie on one side of $S_{1}$ for $s>0$ and jump
260: discontinuously, along the singular line $s=0$, to the other side of $S_{1}$
261: for $s<0$.
262: 
263: \subsection{The Three--Dimensional Horowitz--Polchinski Problem\label{3dHP}}
264: 
265: As we described in the previous section, we want to understand, in general,
266: the gravitational backreaction to a distribution of light--rays in $\mathbb{%
267: M}^{3}\times \mathbb{E}^{D-3}$ distributed on the surface $S_{a}$
268: parametrized in $\mathbb{M}^{3}$ by (\ref{surface}) and fixed at a point
269: along the spectator directions $\mathbb{E}^{D-3}$. For $D\geq 4$ gravity is
270: non--trivial even in the absence of matter, and the problem is not soluble
271: exactly. In $D=3$, on the other hand, space is curved only in the presence
272: of matter and the problem is tractable.
273: 
274: Recall that the basic question is whether black holes are formed as a
275: consequence of the strong gravitational interactions among the light rays.
276: On the other hand, black holes in three dimensional flat space do not exist
277: in the first place, and therefore it seems that the question is not well
278: posed for $D=3$. This is not correct. In fact, the sign of the formation of
279: a black hole in three dimensions is not the presence of a singularity but
280: the existence, in the geometry, of closed timelike curves. To see this,
281: recall again that, in a two--particle scattering with center of mass energy $%
282: \mathcal{E}$ and impact parameter $b$, a black hole forms when 
283: $G\mathcal{E}\gtrsim b^{D-3}$. For $D=3$ the threshold is then 
284: $G\mathcal{E}\gtrsim 1$ and is independent of the impact
285: parameter $b$. To understand the meaning of this threshold, recall that the
286: geometry describing the propagation of a single light ray in 3D is flat
287: everywhere except along a null line, where curvature is concentrated. The
288: local geometry is completely characterized by the holonomy in $SO\left(
289: 1,2\right) $ around the light ray, which is given by a specific null boost 
290: $U_{1}$. In the presence of a second light ray, we have a second holonomy
291: null boost matrix $U_{2}$, as shown in Figure \ref{fig8}. The combined holonomy $%
292: U=U_{1}U_{2}$, on the other hand, depends on the center of mass energy of the two
293: particles. For small center of mass energies $U$ is a rotation. On the other hand, when
294: the center of mass energy exceeds the threshold value \cite{Gott,Deser,Kabat,Gott2}, $U$
295: becomes a boost and closed timelike curves are formed which circle both
296: particles, as shown in Figure \ref{fig8}. Finally, as shown in \cite{Matschull},
297: one may consider the same dynamics in $AdS_{3}$ space, where black holes
298: exist \cite{BTZ}. One then finds that, at exactly the same threshold, the
299: scattering process creates, in the future, a BTZ black hole.
300: 
301: \begin{figure}
302: \begin{center}
303: \includegraphics{fig8.eps}
304: \end{center}
305: \caption{Scattering of light rays in three dimensions. The holonomies $U_1$ 
306: and $U_2$ are null boosts, whereas $U$ is a boost for center of mass energies
307: greater then a specific threshold. In this case, the geometry has CTC's, as
308: indicated. \label{fig8}}
309: \end{figure}
310: 
311: We now see that the three dimensional problem consists in analyzing the
312: gravitational backreaction to light rays located on the surface (\ref
313: {surface}), and investigating the existence, in the full geometry, of
314: closed timelike curves. As we already mentioned, since the surface $S_{a}$
315: preserves the Killing vector $\kappa $, the full geometry with backreaction
316: will also have the same Killing vector. We will show two things:
317: 
318: \begin{enumerate}
319: \item  The geometry will have closed timelike curves. It is important to
320: stress that we are talking about the \textit{uncompactified} geometry, and
321: not of the one obtained after identifying by $e^{\kappa }$.
322: 
323: \item  If we excise from the spacetime manifold the region where $\kappa $
324: is timelike, then the resulting geometry is free from closed timelike
325: curves. As discussed in \cite{CCK}, the excision is necessary to correctly
326: embed these geometries in $M$--theory. This procedure introduces a boundary
327: in the geometry which corresponds, in string/$M$--theory, to the presence of
328: orientifold planes. We will recall these points more at length in the next
329: section.
330: \end{enumerate}
331: 
332: \section{Orientifold versus Orbifold Cosmology\label{orientifold}}
333: 
334: As we just mentioned, and as discussed in detail in \cite{CCK}, when we
335: embed the orbifold spaces in string/$M$--theory we must excise the regions
336: with $\kappa ^{2}<0$. The space $\mathbb{M}^{3}$ is naturally divided into
337: two regions, $\mathcal{X}$ and $\mathcal{Y}$, where the vector $\kappa $ is
338: spacelike (or null) and timelike, respectively. Since $\kappa ^{2}=1+X^{+}X^{-}$, $%
339: \mathcal{X}$ and $\mathcal{Y}$ correspond to the regions $X^{+}X^{-}\geq -1$ and 
340: $X^{+}X^{-}<-1$. We can now consider the quotient manifold 
341: \begin{equation*}
342: \mathbb{M}_{Q}^{3}=\mathbb{M}^{3}/e^{\kappa }=\mathcal{X}_{Q}\cup \mathcal{Y}%
343: _{Q}\, ,
344: \end{equation*}
345: where $\mathcal{X}_{Q}=\mathcal{X}/e^{\kappa }$ and $\mathcal{Y}_{Q}=%
346: \mathcal{Y}/e^{\kappa }$. The action of $\kappa $ is free and therefore the
347: quotient manifolds are smooth. On the other hand, the full quotient space $%
348: \mathbb{M}_{Q}^{3}$, although regular and without boundary, has closed
349: timelike curves which always pass in region $\mathcal{Y}_{Q}$. The space $%
350: \mathcal{X}_{Q}$, on the other hand, does not have closed timelike curves,
351: but does have a boundary.
352: 
353: In the usual time dependent orbifold model, one considers the following
354: compactification of $M$ and $IIA$ theory 
355: \begin{equation*}
356: \begin{array}{ccccc}
357: &  & \mathbb{M}_{Q}^{3}\times \mathbb{T}^{8} &  &  \\ 
358: & \swarrow _{S^{1}} &  & _{\kappa }\searrow  &  \\ 
359: \mathbb{M}_{Q}^{3}\times \mathbb{T}^{7} &  & \overset{TST}{\longleftrightarrow }
360: &  & ?
361: \end{array}
362: \end{equation*}
363: The top line is the $M$--theory vacuum. Going to the left, we are
364: compactifying along a circle $S^{1}$ of the $8$--torus, and we obtain the
365: Type IIA orbifold with constant dilaton. Going to the right, we are
366: compactifying along the Killing vector $\kappa $.
367: The bottom right corner should be related, as usual,
368: by a $TST$ duality transformation to the Type IIA\ orbifold on the bottom
369: left. On the other hand, the IIA supergravity background on the right
370: is quite peculiar, since it has a complex dilaton field. This is 
371: because the vector $\kappa $ is timelike in $\mathcal{Y}_{Q}$.
372: 
373: In \cite{CC, CCK} we have analyzed more carefully the bottom right corner of
374: the diagram and have proposed a related but different compactification of $M$%
375: --theory, which uses crucially non--perturbative objects of the underlying
376: string theory. In fact, it was noted in \cite{CC, CCK} that, if one excises
377: from the $M$--theory vacuum the region $\mathcal{Y}_{Q}$ and dimensionally
378: reduces the space $\mathcal{X}_{Q}\times \mathbb{T}^{8}$ along $\kappa $,
379: one obtains a warped IIA supergravity solution with non--trivial dilaton and
380: RR $1$--form of the form $\mathcal{Z}\times \mathbb{T}^{8}$, where $\mathcal{%
381: Z}$ is a two--dimensional space with boundary. The boundary of $\mathcal{Z}$
382: is singular and corresponds to the supergravity fields of an orientifold
383: plane, which acts as a boundary of spacetime, and has well defined boundary
384: conditions for the string fields. The geometry is best understood after $T$%
385: -dualizing along the $8$--torus and describes then the interaction of a $O8$%
386: --$\overline{O8}$ pair, a system which has also been studied in conformal
387: field theory in \cite{ADSagnotti,Kachru}. We now have a consistent
388: compactification scheme given by 
389: \begin{equation*}
390: \begin{array}{ccccc}
391: &  & \widetilde{\mathcal{X}}_{Q}\times \mathbb{T}^{8} &  &  \\ 
392: & \swarrow _{S^{1}} &  & _{\kappa }\searrow &  \\ 
393: \widetilde{\mathcal{X}}_{Q}\times \mathbb{T}^{7} &  & \overset{TST}{%
394: \longleftrightarrow } &  & \widetilde{\mathcal{Z}}\times \mathbb{T}^{8}
395: \end{array}
396: \end{equation*}
397: where we are adding tildes on the spaces to recall that one has to impose
398: specific boundary conditions on the string fields at the boundaries of these
399: spaces.
400: 
401: \section{Reducing to Two Dimensions\label{mysect}}
402: 
403: We now move back to the analysis of the three--dimensional
404: Horowitz--Polchinski problem. As we already mentioned, the full geometry
405: after backreaction will still have a Killing vector. Therefore, upon
406: dimensional reduction, the problem is effectively two--dimensional.
407: 
408: Let then the general form of the metric in three dimensions be 
409: \begin{equation}
410: ds_{3}^{2}=ds_{2}^{2}+\Phi ^{2}\left( dy+A_{a}dx^{a}\right) ^{2}\,,
411: \label{compactification}
412: \end{equation}
413: where $\frac{\partial }{\partial y}$ is the Killing direction. The 
414: three--dimensional Hilbert action reduces to 
415: \begin{equation*}
416: \int d^{2}x\sqrt{g}\left( \Phi R-\frac{1}{2}\Phi ^{3}F^{2}\right) .
417: \end{equation*}
418: The equation of motion for the field strength $F$ is trivial and implies
419: that the scalar field $\Phi ^{3}\star F$ is constant. By rescaling the
420: variables $y$, $A_{a}$ and $\Phi ^{-1}$ we can fix the constant to any
421: desired value (provided it does not vanish) so that 
422: \begin{equation}
423: \star F=\frac{2}{\Phi ^{3}}.  \label{norm}
424: \end{equation}
425: Then, the equations of motion for the dilaton and the metric can be derived 
426: from the action
427: \begin{equation}
428: \int d^{2}x\sqrt{g}\left( \Phi R-\frac{2}{\Phi ^{3}}\right) .  \label{2Dflat}
429: \end{equation}
430: This action is a particular case of two--dimensional dilaton gravity (see 
431: \cite{Noji,2Dgravity} for reviews and a complete list of references), and we
432: will find it most convenient to work in this framework from now on. Only at
433: the end we will connect back to the discussion in three dimensions.
434: 
435: We conclude this section by showing that (\ref{2Dflat}) is also relevant in
436: the study of the interaction of $O8$--$\overline{O8}$--planes wrapped on a
437: $8$--torus, which again is naturally a
438: two--dimensional problem. In general, the system is described
439: by the following massive IIA background 
440: \begin{eqnarray}
441: E^{2}ds_{10}^{2} &=&\Phi \,ds_{2}^{2}+\Phi ^{-1}\,ds^{2}\left( \mathbb{T}%
442: ^{8}\right) \,,  \label{uplift} \\
443: e^{\phi } &=&g_{s}\Phi ^{-\frac{5}{2}},\,\ \ \ \ \ \ \ \ \ \ \ \ \ \ \star
444: F_{10}=\frac{2E}{g_{s}}\,.  \notag
445: \end{eqnarray}
446: Then the Type IIA action 
447: \begin{equation*}
448: \int d^{10}x\sqrt{G_{10}}e^{-2\phi }\left( R_{10}+4\left( \nabla \phi
449: \right) ^{2}\right) -\frac{1}{2}\int F_{10}\wedge \star F_{10}
450: \end{equation*}
451: becomes the simple two--dimensional dilaton--gravity action 
452: \begin{equation*}
453: \int d^{2}x\sqrt{g}\left( \Phi R-\frac{2 E^{2}}{\Phi ^{3}}\right)\, ,
454: \end{equation*}
455: which again reduces to (\ref{2Dflat}) after rescaling of $\Phi $. Note that,
456: in conformal field theory, we can construct only a single $O8$--$\overline{O8%
457: }$ pair with specific tensions and charges. This then implies (see \cite{CCK}
458: for more details) that 
459: \begin{equation}
460: El_{s}=2g_{s}.  \label{Evsg}
461: \end{equation}
462: 
463: \subsection{Two Dimensional Dilaton Gravity\label{2D}}
464: 
465: In this section we review some basic facts about 2D gravity which will be
466: useful in the analysis of the action (\ref{2Dflat}).
467: 
468: Two--dimensional dilaton gravity is a natural way to define, in two
469: dimensions, theories with a gravitational sector. The basic fields are the
470: matter fields, as well as the metric $g_{ab}$ and a dilaton field $\Phi $,
471: which should be considered together as the gravitational fields.
472: The action has the general form 
473: \begin{equation*}
474: S_{\mathrm{2D}}\left( g,\Phi \right) +S_{\mathrm{M}}\left( g,\Phi ,\mathrm{%
475: Matter}\right) .
476: \end{equation*}
477: $S_{\mathrm{M}}$ is the matter part of the action, whereas $S_{\mathrm{2D}}$
478: is the generalization of the Einstein--Hilbert term and is given by 
479: \begin{equation*}
480: S_{\mathrm{2D}}=\int d^{2}x\sqrt{g}\left[ \Phi R-V\left( \Phi \right) \right]
481: \,,
482: \end{equation*}
483: where $V\left( \Phi \right) $ is a potential for the dilaton. The equations
484: of motion are easily derived to be 
485: \begin{eqnarray}
486: 2\nabla _{a}\nabla _{b}\Phi &=&g_{ab}\left( 2\square \Phi +V\right) -\tau
487: _{ab}\,\ ,  \label{metric-eq} \\
488: R &=&\frac{dV}{d\Phi }+\rho \,,  \notag
489: \end{eqnarray}
490: where $\tau _{ab}$ and $\rho $ are 
491: \begin{equation*}
492: \tau _{ab}=-\frac{2}{\sqrt{g}}\frac{\delta S_{\mathrm{M}}}{\delta g^{ab}}%
493: \,,\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \rho =-\frac{1%
494: }{\sqrt{g}}\frac{\delta S_{\mathrm{M}}}{\delta \Phi }.
495: \end{equation*}
496: Moreover, the conservation of the stress--energy tensor $\tau _{ab}$ is
497: modified by the dilaton current $\rho $ to 
498: \begin{equation*}
499: \nabla ^{a}\tau _{ab}+\rho \nabla _{b}\Phi =0.
500: \end{equation*}
501: 
502: The inherent simplicity of the dilaton gravity model lies in the following
503: observations \cite{Kunstatter, Noji, 2Dgravity}. Define $J\left( \Phi \right) $ by 
504: \begin{equation*}
505: J=\int Vd\Phi 
506: \end{equation*}
507: and consider the function 
508: \begin{equation}
509: C=\left( \nabla \Phi \right) ^{2}+J\left( \Phi \right) 
510: \label{Cfunct}\end{equation}
511: and the vector field 
512: \begin{equation*}
513: \kappa ^{a}=\frac{2}{\sqrt{g}}\epsilon ^{ab}\nabla _{b}\Phi \,.
514: \end{equation*}
515: Then, whenever $\tau _{ab}=\rho =0$ -- \textit{i.e. }for any \textit{vacuum}
516: solution to the equations of motion -- the function $C$ is constant and $%
517: \kappa $ is a Killing vector of the solution. The first fact follows
518: immediately from the equations of motion which imply
519: \begin{equation}
520: \nabla _{a}C=-\tau _{ab}\nabla ^{b}\Phi +\nabla _{a}\Phi \,(\tau
521: _{bc}g^{bc})\,.  \label{Cvariation}
522: \end{equation}
523: The second fact is proved most easily in conformal coordinates $z^{\pm }$,
524: with metric 
525: \begin{equation*}
526: -dz^{+}dz^{-}e^{\Omega }\,.
527: \end{equation*}
528: Then we have that 
529: \begin{equation*}
530: \kappa _{\pm }=\mp \nabla _{\pm }\Phi 
531: \end{equation*}
532: and the non--trivial Killing equations become $\nabla _{+}\nabla _{+}\Phi
533: =\nabla _{-}\nabla _{-}\Phi =0$, which now follow from (\ref{metric-eq})
534: whenever $\tau _{ab}=0$. Finally note that these equations are equivalent to 
535: \begin{equation*}
536: \partial _{-}\kappa ^{+}=\partial _{+}\kappa ^{-}=0\,.
537: \end{equation*}
538: 
539: Using these facts it is trivial to find all classical vacuum solutions. In
540: fact, whenever $\kappa ^{2}<0$, the metric can be put locally into the form 
541: \begin{equation}
542: ds^{2}=-f\left( x\right) dt^{2}+dx^{2}\,,\,\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
543: \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \kappa =\alpha \frac{\partial }{\partial t}%
544: \,,  \label{sol1}
545: \end{equation}
546: where $\alpha $ is a constant. Then, one has the explicit solution of the
547: equations of motion 
548: \begin{equation}
549: C=\left( \frac{d\Phi }{dx}\right) ^{2}+J\left( \Phi \right) \,,\ \ \ \ \ \ \
550: \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ f=\frac{4}{\alpha ^{2}}\left( 
551: \frac{d\Phi }{dx}\right) ^{2}\,,  \label{sol2}
552: \end{equation}
553: which depends only on the constant $C$. Similar equations hold in regions
554: where $\kappa ^{2}>0$.
555: 
556: Let us now analyze the geometry in the presence of matter. For our purposes,
557: we are going to consider only matter lagrangians $S_{M}$ which do not depend
558: on the dilaton, and which are conformal. This implies that 
559: \begin{eqnarray*}
560: &&\tau _{+-} =\rho =0 \,, \\
561: &&\partial _{-}\tau _{++} = \partial _{+}\tau _{--}=0\,.
562: \end{eqnarray*}
563: The simplest example is clearly a conformally coupled scalar $\eta $ with action $%
564: S_{M}=-\int \left( \nabla \eta \right) ^{2}$. The effect of this type of
565: matter is best described by considering a shock wave \cite{CGHS,Lawrence},
566: which is represented in conformal coordinates by a stress--energy tensor of
567: the form 
568: \begin{equation*}
569: \tau _{--}\left( z^{-}\right) =\epsilon \,\delta \left(
570: z^{-}-z_{0}^{-}\right) \,.\,\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
571: \left( \epsilon >0\right) 
572: \end{equation*}
573: \begin{figure}
574: \begin{center}
575: \includegraphics{fig6.eps}
576: \end{center}
577: \caption{Shock wave solution in two--dimensional dilaton gravity.\label{fig6}}
578: \end{figure}
579: The positivity of $\epsilon $ can be understood by looking at the
580: conformally coupled scalar, for which $\tau _{--}=2\left( \nabla _{-}\eta
581: \right) ^{2}>0$. Recalling from (\ref{Cvariation}) that 
582: \begin{equation}
583: \nabla _{-}C=2\tau _{--}\nabla _{+}\Phi \,e^{-\Omega }=\tau _{--}\kappa
584: ^{-}\,,  \label{Cjump}
585: \end{equation}
586: we conclude that the shock front interpolates, as we move along $z^{-}$,
587: between the vacuum solution with $C=C_{0}$ and the vacuum solution with $%
588: C=C_{0}+\epsilon \kappa ^{-}\left( z_{0}^{-}\right) $ (see Figure \ref{fig6}). As a
589: consistency check note that, since in the vacuum $\tau _{--}$ and $\kappa
590: ^{-}$ are functions only of $z^{-}$, equation (\ref{Cjump}) defines a jump
591: in the function $C$ which is independent of the position $z^{+}$ along the
592: shock wave.
593: 
594: 
595: \subsection{The Case $V=2\Phi ^{-3}$}
596: 
597: We now specialize to the action (\ref{2Dflat}) by taking 
598: \begin{equation*}
599: V=\frac{2}{\Phi ^{3}}\,,\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
600: \ \ \ \ J=-\frac{1}{\Phi ^{2}}\,.
601: \end{equation*}
602: First consider the vacuum solutions. These are easy to obtain by considering
603: the flat metric (\ref{flat}) written in the coordinates $x^{\pm
604: }=X^{\pm }e^{\mp EY}$ and $y=\sqrt{E}Y$, where $E>0$ is a constant with
605: units of energy, 
606: \begin{equation*}
607: \frac{\partial }{\partial y}=\sqrt{E}\left( X^{+}\partial _{+}-X^{-}\partial
608: _{-}\right) +\frac{1}{\sqrt{E}}\,\partial _{Y}
609: \end{equation*}
610: is a Killing direction and $x^{\pm }$ are the two--dimensional coordinates.
611: If one rewrites the metric in the form (\ref{compactification}), it is easy
612: to check the normalization (\ref{norm}) and to compute the two--dimensional
613: metric and dilaton field. The explicit computation is done in \cite{CC} and
614: the result corresponds to the cosmological solution shown in Figure 
615: \ref{fig1}, where the timelike orientifold singularities
616: are at a distance of order $E^{-1}$. For future use, we divide the diagram
617: in various regions I$_{in,out}$, II$_{L,R}$ and III$_{L,R}$, as shown in the
618: Figure.
619: 
620: \begin{figure}
621: \begin{center}
622: \includegraphics{fig1.eps}
623: \end{center}
624: \caption{The cosmological vacuum solution for $V=2\Phi^{-3}$. Regions I$_{in}$, 
625: I$_{out}$ are the contracting and expanding cosmologies, whereas regions II$_{L,R}$
626: are the intermediate regions. Regions III$_{L,R}$ are after the singularities, and 
627: correspond, when uplifted to three dimensions, to the regions where the
628: Killing vector $\kappa$ becomes timelike.\label{fig1}}
629: \end{figure}
630: 
631: Analytic results are simplest in Lorentzian polar coordinates. For example,
632: in the Rindler wedges II$_{L}$ and II$_{R}$, where $x^{+}x^{-}<0$, one uses
633: coordinates $t,x$ defined by $x^{\pm }=\pm xe^{\pm t}$ and obtains the static
634: solution \cite{CC} 
635: \begin{eqnarray}
636: ds_{2}^{2} &=&-dt^{2}\left( \frac{x^{2}}{1-E^{2}x^{2}}\right) +dx^{2}\,,
637: \label{vacuumsol} \\
638: \Phi  &=&\frac{1}{\sqrt{E}}\sqrt{1-E^{2}x^{2}}\,.  \notag
639: \end{eqnarray}
640: It is easy to check that the above solution satisfies (\ref{sol1}) and (\ref
641: {sol2}) with 
642: \begin{equation*}
643: C=-E.
644: \end{equation*}
645: As we already discussed in section \ref{orientifold}, these $C<0$ solutions
646: correspond to non BPS $O8$--$\overline{O8}$ geometries when uplifted to
647: string theory with equation (\ref{uplift}). Notice that, given 
648: $E$\thinspace (or equivalently $g_{s}$ by equation (\ref{Evsg}%
649: )), one has no freedom in the solution, which is unique. Therefore, \textit{%
650: no fine tuning is required in the initial conditions for the metric and the
651: dilaton in order to obtain a solution with a bounce and with past and
652: future cosmological horizons.} Recall also that the conformal field theory description
653: of the $O\overline O$ system has two distinct moduli, which are the string coupling and
654: the distance between the orientifolds in string units. We therefore see that
655: \textit{the backreaction of the geometry ties the two moduli with equation (\ref{Evsg})
656: and reduces them to a single free parameter.}
657: 
658: We also notice that the $C=0$ solution is clearly singled out and corresponds to 
659: the supergravity solution of a single BPS $O8$--plane.
660: The $C>0$ solutions do not have a clear interpretation in string
661: theory, and it would be nice to understand this point further. 
662: 
663: Before considering more general solutions in the presence of matter, let us
664: discuss the coordinate change from $x^{\pm }$ to conformal coordinates $%
665: z^{\pm }$. For the sake of simplicity, and since this is all we will need,
666: we write formulae for the case $E=1$. We consider then the change of
667: coordinates 
668: \begin{equation}
669: z^{+}z^{-}=e^{2\Phi }\,\frac{\Phi -1}{\Phi +1},\,\ \ \ \ \ \ \ \ \ \ \ \ \ \
670: \ \ \ \ \ \ \ \ \ \ \frac{z^{+}}{z^{-}}=\frac{x^{+}}{x^{-}},
671: \label{confcoord}
672: \end{equation}
673: where 
674: \begin{equation*}
675: \Phi =\sqrt{x^{+}x^{-}+1}.
676: \end{equation*}
677: Then, it is simple to show that the metric becomes conformal 
678: \begin{equation}
679: ds_{2}^{2}=-\frac{dz^{+}dz^{-}}{z^{+}z^{-}}\left( \frac{\Phi ^{2}-1}{\Phi
680: ^{2}}\right) .
681: \label{confmetric}\end{equation}
682: 
683: Finally let us add matter and consider shock wave solutions. Given the
684: discussion in section (\ref{2D}), it is now almost trivial to find the
685: correct geometries, which are given pictorially in Figure \ref{fig3}. In Figure
686: \ref{fig3}a, before the shock wave, we have the vacuum solution with some given
687: value of $E$. After the wave, one has again a vacuum solution, but with a
688: different value $E^{\prime }$. Recall that 
689: \begin{equation*}
690: E^{\prime }=E-\epsilon \kappa ^{-}
691: \end{equation*}
692: where $\epsilon >$ $0$ and $\kappa ^{-}=2\nabla _{+}\Phi \,e^{-\Omega }$
693: must be computed along the wave. As we move in Figure \ref{fig3}a from point $a$ to
694: points $b$ and $c$ along the shock wave, in the direction of increasing $%
695: z^{+}$, the value of $\Phi $ decreases to $0$ at $c$ on the singularity.
696: Therefore $\nabla _{+}\Phi \,<0$ and one has that 
697: \begin{equation*}
698: E^{\prime }>E.
699: \end{equation*}
700: Moreover, in any vacuum solution with $C=-E$, the value of the
701: dilaton on the horizons is $1/\sqrt{E}$, as can
702: be seen from equation (\ref{vacuumsol}) at $x=0$.
703: Therefore, since the value of $\Phi $ is continuous across the shock wave,
704: we notice that the horizon to the left of the wave, where the dilaton has
705: value $\Phi =1/\sqrt{E^{\prime }}$, must intersect the wave between\ the
706: points $b$ and $c$, as drawn.
707: 
708: \begin{figure}
709: \begin{center}
710: \includegraphics{fig3.eps}
711: \end{center}
712: \caption{Shock wave solutions in the cosmological geometry.\label{fig3}}
713: \end{figure}
714: 
715: A particular case of the shock wave geometry is attained when the wave moves
716: along one of the horizons, as shown in Figure \ref{fig3}b. Since $\Phi $ is constant
717: along the horizons, we have that $\nabla _{+}\Phi =0$ and therefore that 
718: \begin{equation*}
719: E^{\prime }=E.
720: \end{equation*}
721: Let us breifly explain why the horizon at a constant value of $z^+$ shifts 
722: as one passes the shock wave. It is easy to see that the horizon in question is given by
723: the curve $\kappa^+ = 0$. In the vacuum, $\kappa^+$ is a function of $z^+$ alone,
724: but in the presence of matter one has that
725: \begin{equation*}
726: \partial_- \kappa^+ = e^{-\Omega} \tau_{--}\,.
727: \end{equation*}
728: Then, since $\Omega$ is constant along the horizons (and therefore along the shock
729: wave) and since $\tau_{--}$ has a delta singularity, the function $\kappa^+$
730: just jumps by a finite constant across the wave, thus explaining the shift in
731: the position of the horizon.
732: 
733: Let us conclude by noticing that the instability of the null orbifold found in \cite{Lawrence}
734: is again a shock wave which interpolates between two different vacuum solutions. 
735: The initial solution is a BPS vacuum ($C=0$), which has a different spacetime structure
736: from the other solutions with $C\neq0$ which are formed after the shock wave.
737: In our case this instability does not arise since we are interpolating between
738: two non--BPS vacua with the same global structure.
739: 
740: \subsection{The Shock Wave Solution in Three Dimensions}
741: 
742: The geometry of Figure \ref{fig3}a corresponds, when uplifted to three dimensions,
743: to the gravitational backreaction to matter distributed along a specific
744: surface, which is the uplift of the shock wave curve. In order to better
745: understand the geometry of this surface, consider first the shock wave
746: without backreaction. From now on we take, for simplicity, $E=1$ and $%
747: E^{\prime }=1+\delta $ for some $\delta >0$. The limiting geometry for $%
748: \delta \rightarrow 0$ is given by Figure \ref{fig2}. The underlying 2D geometry is
749: the vacuum $E=1$ solution, and we have singled out a specific curve, which
750: is given, in the conformal coordinates $z^{\pm }$, by the two segments 
751: \begin{eqnarray}
752: z^{-} &=&-1\,,\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ (-\infty <z^{+}<1)
753: \label{cur} \\
754: z^{+} &=&1\,.\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
755: (-1<z^{-}<\infty )  \notag
756: \end{eqnarray}
757: 
758: \begin{figure}
759: \begin{center}
760: \includegraphics{fig2.eps}
761: \end{center}
762: \caption{The shock wave before backreaction, both in conformal coordinates
763: $z^\pm$ and in the coordinates $x^\pm$ of equation (\ref{confcoord}).\label{fig2}}
764: \end{figure}
765: 
766: We now uplift this configuration to three dimensions. The vacuum solution
767: becomes, as we emphasized before, nothing but flat space. The curve
768: becomes then a specific surface in $\mathbb{M}^{3}$, whose geometry is most
769: easily analyzed by passing to the coordinates $x^{\pm }$ in (\ref{confcoord}%
770: ). It is simple to see, using the explicit coordinate transformation (\ref
771: {confcoord}), that the curve (\ref{cur}) becomes (see Figure \ref{fig2}) 
772: \begin{equation}
773: x^{\pm }=\left( s\pm 1\right) e^{\mp s}  \label{cur2}
774: \end{equation}
775: for $s\in \mathbb{R}$. The first part of the curve (\ref{cur}) corresponds
776: to (\ref{cur2}) for $s<0$, whereas the second part corresponds to $s>0$.
777: Comparing with (\ref{surxpxm}) we immediately see that this curve
778: corresponds, when uplifted to three dimensions, to the Horowitz--Polchinski
779: surface $S_{1}$. Recall from section \ref{HPproblem} that, among the various
780: surfaces $S_{a}$, the surface $S_{1}$ is special since it is the only null
781: surface, and it is therefore not surprising that we are obtaining exactly $%
782: S_{1}$ from the two--dimensional shock wave.
783: 
784: We therefore conclude that \textit{the geometry of Figure \ref{fig3}a, when uplifted
785: to three dimensions, represents the gravitational backreaction to light rays
786: distributed on the HP surface }$S_{1}$\textit{.}
787: 
788: \subsection{Finding the Closed Timelike Curve}
789: 
790: 
791: We are now in a position to investigate the existence of closed timelike
792: curves in the geometry of Figure \ref{fig3}a. Recall that, without shock wave,
793: the quotient geometry has closed timelike curves that always pass
794: through region III, and that are removed by excising this region. These
795: curves are \textit{not closed in the covering space}, which is flat. In the
796: geometry with the shock wave, we shall see that there are closed timelike
797: curves \textit{already in the covering space}, which could signal an instability. However
798: these curves always pass through region III, and again should be excised.
799: 
800: In order to describe the closed timelike curves, we first need to prove a simple fact.
801: Consider, in $\mathbb{M}^{3}$, two points $A$ and $B$ which are connected by
802: a future--directed timelike curve from $A$ to $B$. Let $x_{A}^{\pm },0$ and $%
803: x_{B}^{\pm },y_{B}$ be the coordinates of these two points in the coordinate
804: system (\ref{2dcoord}) and let us assume that both points are in Region II$%
805: _{R}$, \textit{i.e.} the region with $x^{+}>0$, $x^{-}<0$ and $x^{+}x^{-}<1
806: $. We parametrize the path from $A$ to $B$ as $x^{\pm }\left( s\right)
807: ,y\left( s\right) $, with $0\leq s\leq 1$. First of all, it is easy to show
808: that the full path must be entirely regions II$_{R}$ or III$_{R}$%
809: , because otherwise the horizons in the geometry prevent the curve from
810: returning to region II$_{R}$ without becoming spacelike. Therefore, since we
811: are confined to the right Rindler wedge of the $x^{\pm }$ plane, we may
812: adopt polar coordinates $x^{\pm }=\pm x\,e^{\pm t}$. We want to show that, 
813: \textit{if }$t_{A}>t_{B}$\textit{\ (the point }$A$\textit{\ is after the
814: point }$B$\textit{\ in Rindler time), then any forward directed timelike
815: curve from }$A$\textit{\ to }$B$\textit{\ must go into region III}$_{R}$ 
816: \textit{(the region with }$x>1$\textit{). }To prove this fact, recall first
817: the metric on $\mathbb{M}^{3}$ in the coordinates $x,t,y$%
818: \begin{equation*}
819: dx^{2}-x^{2}dt^{2}+dy^{2}\left( 1-x^{2}\right) -2x^{2}dydt\,.
820: \end{equation*}
821: Consider then the function $t\left( s\right) $. It starts at $t\left(
822: 0\right) =t_{A}$ and it is increasing for small values of $s$. This is
823: because surfaces of constant $t$ are spacelike in region II$_{R}$ (since $%
824: dx^{2}+dy^{2}\left( 1-x^{2}\right) >0$ if $x^{2}<1$) and therefore future
825: directed timelike curves have always $\frac{dt}{ds}>0$ in region II$_{R}$.
826: Now, since by assumption $t_{B}<t_{A}$, the function $t\left( s\right) $ must
827: achieve a maximum for some value of $s$. Then, using the same reasoning, the
828: curve must be in region III$_{R}$ to be timelike, and this concludes the
829: proof. We can actually always construct a future directed timelike curve by
830: connecting the two points $A$ and $B$ with a straight line, provided that we
831: choose $y_{B}$ large enough, as can be seen by direct computation.
832: 
833: We can now prove the existence of a closed timelike curve in the full geometry
834: (\ref{fig3}a). The regions to the left and to the right of the wave are vacuum 
835: solutions which correspond, when uplifted to three dimensions, to regions in 
836: flat Minkowski space. We can then think of the full geometry as follows. Let 
837: $\widetilde{\mathbb{M}}^{3}$ and $\mathbb{M}^{3}$ be two copies of three--dimensional
838: space which are going to describe the geometry to the left and to the right
839: of the shock wave, and which we parametrize with coordinates
840: \begin{eqnarray*}
841: &&\widetilde{X}^\pm=\widetilde{x}^{\pm }e^{\pm \sqrt{E^{\prime }}\widetilde{y}}\,\ \ \ \ \ \ \
842: \ \
843: \widetilde{Y}=\frac{1}{\sqrt{E^{\prime }}}\ \widetilde{y}\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
844: \ \ \ \ \ \ (\widetilde{\mathbb{M}}^{3}) \\
845: &&X^\pm=x^{\pm }e^{y}\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ Y=y\,\ \ \ \ \ \ \ \ \ \ \ \ \
846: \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ (\mathbb{M}^{3})\
847: \end{eqnarray*}
848: respectively. The shock wave itself defines
849: surfaces $\widetilde{S}$ and $S$ in the two spaces $\widetilde{\mathbb{M}}^{3}$
850: and $\mathbb{M}^{3}$. The full geometry is then obtained as follows. Denote
851: the region to the left of $\widetilde{S}$ in $\widetilde{\mathbb{M}}^{3}$ 
852: by $\widetilde{\mathbb{M}}^{3}_L$. This corresponds to the part of the 
853: geometry to the left of the shock wave. Similarly, denote 
854: the region to the right of ${S}$ in ${\mathbb{M}}^{3}$ by ${\mathbb{M}}^{3}_R$,
855: which is the region to the right of the wave. The full geometry is then 
856: given by taking the two regions $\widetilde{\mathbb{M}}^{3}_L$ and
857: ${\mathbb{M}}^{3}_R$ and gluing the two boundaries $\widetilde{S}$
858: and ${S}$ in a specific way. We do not need the details of the gluing, aside
859: from the simple fact that $\widetilde{y} = y$ when connecting $\widetilde{S}$
860: and ${S}$. 
861: \begin{figure}
862: \begin{center}
863: \includegraphics{fig4.eps}
864: \end{center}
865: \caption{A closed timelike curve in the geometry induced by the surface $S_1$.\label{fig4}}
866: \end{figure}
867: Looking at Figure \ref{fig4}, which represents the shock wave solution 
868: in conformal coordinates, choose then the two points $A$ and $B$ as drawn.
869: They will correspond to points on the surface $\widetilde{S}$ with coordinates 
870: $\tilde{x}_{A}^{\pm },\tilde{y}_{A}$ and $\tilde{x}_{B}^{\pm },\tilde{y}_{B}$ and
871: on $S$ with coordinates ${x}_{A}^{\pm },y_A$ and ${x}_{B}^{\pm },{y}_{B}$,
872: where, as we already noticed, one has $\tilde{y}_{A}={y}_{A}$ and $\tilde{y}_{B}={y}_{B}$.
873: As in the discussion in the previous part of this section, we choose $y_{A}=0$ and $%
874: y_{B}$ large and positive.
875: Notice first that, with respect to the geometry ${\mathbb{M}}^{3}_R$ on the right of the
876: shock wave, both points $A$ and $B$ are in region II$_{R}$, whereas they are in 
877: regions I$_{out}$ and I$_{in}$ relative to the geometry $\widetilde{\mathbb{M}}^{3}_L$ 
878: on the left of the wave. In equations, we have that
879: \begin{eqnarray*}
880: x_{A}^{+} &>&0,\,\ \ \ \ \ \ \ \ \ x_{A}^{-}<0,\,\ \ \ \ \ \ \ \ \
881: x_{A}^{+}x_{A}^{-}<1, \\
882: x_{B}^{+} &>&0,\,\ \ \ \ \ \ \ \ \ x_{B}^{-}<0,\,\ \ \ \ \ \ \ \ \
883: x_{B}^{+}x_{B}^{-}<1,\,
884: \end{eqnarray*}
885: and
886: \begin{eqnarray*}
887: \widetilde{x}_{A}^{\pm } &>&0, \\
888: \widetilde{x}_{B}^{\pm } &<&0.
889: \end{eqnarray*}
890: This is the key feature which is required in order to have a closed timelike curve. 
891: Let us first concentrate on the right part of the wave. We may just consider the 
892: straight line from $A$ to $B$ in ${\mathbb{M}}^{3}_R$ which is
893: future directed for $y_{B}$ large, and which goes, when projected onto the $%
894: z^{\pm }$ plane as in Figure \ref{fig4}, in the region after the singularity. This
895: is because the point $A$ is after point $B$ in the Rindler time of region II$%
896: _{R}$ ($x^+_B/x^-_B>x^+_A/x^-_A$), and so we can apply the arguments of the first part of the section.
897: Now we consider the region to the left of the shock wave. It is easy to
898: check, by direct computation, that the straight segment from $B$ to $A$ in 
899: $\widetilde{\mathbb{M}}^{3}_L$ is still timelike for large values of 
900: $\tilde{y}_{B}=y_B$, and it is clearly future directed since $B$ is in 
901: I$_{in}$ and $A$ is in I$_{out}$. This closes the
902: loop, as drawn in Figure \ref{fig4}, and we have found a closed timelike curve.
903: 
904: Note that, due to the simple lemma proved above, it must be that our closed
905: curve always goes into region III. Recall though that the boundary of
906: region III corresponds in string theory to the orientifold $O8$--plane, and
907: therefore region III should be excised, as described in section \ref
908: {orientifold}. \textit{The final }$M$\textit{--theory geometry is free of
909: closed timelike curves, and we have then shown that the presence of the
910: orientifolds not only resolves the issues put forward in section \ref
911: {orientifold}, but also cures the instability due to the formation of large
912: black holes.}
913: 
914: \subsection{Discussion}
915: 
916: Let us conclude this section with some comments.
917: 
918: First of all, we can try to consider the geometry induced by light rays on
919: surfaces $S_{a}$ for $-1<a<1$ (so that the surface does not go into region
920: III). This corresponds, in 2D gravity, to matter which couples to the
921: dilaton and to the conformal factor of the metric, and is therefore
922: analytically more complex. Nonetheless, we can understand pictorially, in
923: Figure \ref{fig5}, that the physics is qualitatively unaltered. In particular
924: one can find, as before, closed timelike curves which necessarily have to
925: pass in region III.
926: 
927: \begin{figure}
928: \begin{center}
929: \includegraphics{fig5.eps}
930: \end{center}
931: \caption{A closed timelike curve in the geometry induced by the surface $S_a$
932: for $a<1$.\label{fig5}}
933: \end{figure}
934: 
935: Secondly, it is clear that, in Figure \ref{fig3}a, it is crucial that $%
936: E^{\prime }>E$. This fact, recall, comes from positivity requirements on the
937: stress--energy tensor $\tau _{ab}$. If $E^{\prime }$ had been less then $E$,
938: the graph would have had a different structure. In particular, the horizon
939: of the left part of the geometry would have intersected the shock wave not
940: between the points $b$ and $c$, but between $a$ and $b$. The causal
941: structure of the space, when uplifted to three dimensions, is then quite
942: different and it is easy to show, using arguments similar to those in the
943: previous section, that the 3D geometry has no closed timelike curves.
944: 
945: Finally, let us comment on the general Horowitz--Polchinski problem for $D>3$.
946: In this case, gravity is non--trivial even in the absence of matter, and therefore
947: an exact solution to the problem is probably out of reach. On the other
948: hand, it seems unlikely that a correct guess on the final qualitative features of the 
949: solution can be obtained by looking at the interaction between two (or, for that matter,
950: a finite number) of light--rays. This fact is already true if we just consider the 
951: \textit{linear reaction} of the gravitational field to the matter distributed
952: on the HP surfaces $S_a$. Then, very much like in electromagnetism, it is incorrect 
953: to guess the qualitative features of fields by looking at just a finite subset
954: of the charges (matter in this case), whenever the charge distribution is infinite
955: (this infinity is really not an approximation in this case, since it comes
956: from the infinite extent of the surface $S_a$ due to the unwrapping of the 
957: quotient space). Therefore, to decide if the problem exists in higher dimensions, 
958: much more work is required, already in the linear regime of gravity, but most
959: importantly in the full non--linear setting. Note that \textit{the only case
960: in which the HP argument is fully correct is exactly in dimension $D=3$, where
961: the gravitational interaction is topological and when, therefore, the interaction
962: of an infinite number of charges can be consistently analyzed by breaking it down into 
963: finite subsets}. This indeed is what we find in the previous sections, since 
964: closed timelike curves do appear.
965: 
966: 
967: \section{Stability of the Cosmological Cauchy Horizon}
968: 
969: Another related classical instability which can arise in orbifold and orientifold
970: cosmologies is due to the backreaction of general matter fields as they
971: propagate through the bounce: as the universe contracts, particles will
972: accelerate and will create a large backreaction in the geometry. General
973: instabilities of this type can only be studied in the linearized regime, as
974: opposed to the shock wave geometries which could be analyzed fully,
975: including gravitational backreaction. In fact, in the previous section we concentrated on
976: conformally coupled non--dilatonic matter propagating in the
977: cosmological geometry, and this was possible due to the simple coupling to the
978: two--dimensional gravitational fields $\Phi $ and $g_{ab}$. We now want to
979: consider more general matter fields, either non--conformal or
980: coupled to the dilaton. In the following, we shall concentrate on a scalar
981: propagating in the quotient space $\mathbb{M}^{3}/e^{\kappa }$, which
982: corresponds to a free massive scalar on the covering space $\mathbb{M}^{3}$.
983: Our results will apply to the orientifold cosmology after imposing
984: Dirichlet or Neumann boundary conditions on the surface $X^{+}X^{-}=-E^{2}$%
985: . These wave functions were studied in \cite{CCK}, and we shall review
986: briefly those results before considering the issue of stability.
987: 
988: \subsection{Particle States}
989: 
990: Let us start with a massive field $\Psi $ of mass $m$ on the covering space $%
991: \mathbb{M}^{3}$, which satisfies the Klein--Gordon equation 
992: \begin{equation}
993: \square \,\Psi (X)=m^{2}\,\Psi (X)\ .
994: \label{KG}\end{equation}
995: We demand that $\Psi $ be invariant under the orbifold action, so that 
996: \begin{equation}
997: \Psi (X)=\Psi (e^{\kappa }X)\ ,  \label{bc}
998: \end{equation}
999: where $\kappa =2\pi i\left( \Delta J+RK\right) $ and 
1000: \begin{eqnarray*}
1001: iJ &=&X^{+}\partial _{+}-X^{-}\partial _{-}\ , \\
1002: iK &=&\partial _{Y}\ .
1003: \end{eqnarray*}
1004: The energy scale $E$ of section \ref{mysect} is given by $E=\Delta R^{-1}$.
1005: Since the quotient space has the symmetry generated by the Killing vector $%
1006: \kappa $, and since the operators $J$, $K$ and $\square $ commute, it is
1007: convenient to choose a basis of solutions to (\ref{KG}) and (\ref{bc})
1008: where the operators $J$ and $K$
1009: are diagonal. We then choose our field $\Psi $ so that 
1010: \begin{eqnarray*}
1011: J\,\Psi (X) &=&p\,\Psi (X)\ , \\
1012: K\,\Psi (X) &=&k\,\Psi (X)\ ,
1013: \end{eqnarray*}
1014: by writing 
1015: \begin{equation*}
1016: \Psi _{p,k}\left( X\right) =\Psi _{p}(X^{+},X^{-})\, e^{ikY} \,.
1017: \end{equation*}
1018: The wave functions $\Psi _{p}$ are functions of Lorentzian spin $J=p$ which also
1019: satisfy the two--dimensional Klein--Gordon equation $\left( 4\partial
1020: _{+}\partial _{-}+\omega ^{2}\right) \Psi _{p}=0$, where $\omega
1021: ^{2}=m^{2}+k^{2}$. They can then be solved in Lorentzian polar coordinates,
1022: as we shall review below, in terms of Bessel functions, which naturally have
1023: an integral representation as a sum of the standard plane waves in the
1024: covering space. In terms of the quantum numbers $p$, $k$, the orbifold
1025: boundary condition (\ref{bc}) becomes simply 
1026: \begin{equation}
1027: \Delta \,p+\,R\,k=n\in \mathbb{Z\,}.  \label{quant}
1028: \end{equation}
1029: Generally, when $R\neq 0$, the spectrum of $p\in {\mathbb{R}}$ is continuous
1030: for each $n$. On the other hand, for the pure boost orbifold ($R=0$), the
1031: spectrum is discrete since $\Delta \,p=n$. This difference is crucial. \ We
1032: shall see more in detail in the next section that, in the general case, 
1033: \textit{it is possible to define particle states that do not destabilize the
1034: cosmological vacuum on the horizons by properly integrating over the various
1035: spins }$p$, whereas in the pure boost case there is an infinite blue shift
1036: as particles approach the tip of the Minkowski cone \cite{HS}. The situation
1037: is analogous to particles propagating in the ``null boost'' and ``null boost 
1038: $+$ translation'' orbifolds studied in \cite{LMS2}. From now on we consider
1039: the general case $R\neq 0$, and we use as quantum numbers $p$ and $n$, thus
1040: labeling the wave functions $\Psi _{p,n}$. 
1041: 
1042: To study the behaviour of the wave functions describing the collapse of
1043: matter through the contracting region I$_{in}$ we first move to light--cone coordinates $%
1044: x^{\pm }=X^{\pm }e^{\mp EY}$ and then to polar
1045: coordinates in the Milne wedge of the $x^{\pm }$ plane $x^{\pm }=t\,e^{\pm Ex}$. This means
1046: that, using the quantization condition (\ref{quant}), the wave function $%
1047: \Psi _{p,n}$ is given by 
1048: \begin{equation}
1049: \Psi _{p,n}\left( X\right) =\Psi _{p}^{\left( \pm \right) }(x^{+},x^{-})
1050: \, e^{i\frac{n}{R}Y} \,,  \label{basicsol}
1051: \end{equation}
1052: where the functions $\Psi _{p}^{\left( \pm \right) }$ have the form 
1053: \begin{equation*}
1054: \Psi _{p}^{\left( \pm \right) }=J_{\pm ip}(\omega |t|)\,e^{ipEx}
1055: \end{equation*}
1056: and where $J_{\pm ip}$ are the Bessel functions of imaginary order $%
1057: \pm ip$. In terms of the quantum numbers $p$ and $n$, the frequency $\omega $
1058: satisfies the mass shell condition 
1059: \begin{equation*}
1060: \omega ^{2}=m^{2}+\left( Ep-\frac{n}{R}\right) ^{2}\ .
1061: \end{equation*}
1062: 
1063: \subsection{Near Horizon Behaviour of the Wave Functions}
1064: 
1065: We are interested in the limit $t\rightarrow 0$ near the cosmological
1066: horizon, so it is convenient to recall the basic functional form of the
1067: Bessel functions 
1068: \begin{equation*}
1069: J_{\pm ip}(\omega |t|)=\left( \frac{\omega |t|}{2}\right) ^{\pm ip}F_{\pm
1070: ip}\left( \omega ^{2}t^{2}\right) \ ,
1071: \end{equation*}
1072: where $F_{\pm ip}$ is an entire function on the complex plane, which has
1073: the expansion 
1074: \begin{equation*}
1075: F_{\pm ip}(z)=\sum_{k=0}^{\infty }\frac{(-)^{k}}{4^{k}\,k!\,\Gamma (k+1\pm
1076: ip)}\,z^{k}\ .
1077: \end{equation*}
1078: Then, in terms of the two--dimensional light--cone coordinates $x^{\pm }$,
1079: which are well defined throughout the whole geometry, we have 
1080: \begin{equation*}
1081: \Psi _{p}^{\left( \pm \right) }=\left( \frac{\omega |x^{\pm }|}{2}\right)
1082: ^{\pm ip}F_{\pm ip}\left( \omega ^{2}x^{+}x^{-}\right) \ .
1083: \end{equation*}
1084: These wave functions are well behaved everywhere except at the horizons $%
1085: x^{\pm }=0$, where there is an infinite blue--shift of the frequency. To see
1086: this, consider the leading behaviour of the wave function $\Psi _{p}^{(+)}$
1087: at both horizons ($x^{+}=0$ or $x^{-}=0$) 
1088: \begin{equation*}
1089: \Psi _{p}^{(+)}\sim \left( \frac{\omega |x^{+}|}{2}\right) ^{ip}\frac{1}{%
1090: \Gamma (1+ip)}\ .
1091: \end{equation*}
1092: Near the horizon $x^{-}=0$ the wave function is well behaved and can be
1093: trivially continued through the horizon. Near $x^{+}=0$, on the other hand,
1094: the wave function has a singularity which can be problematic. In fact, close
1095: to the horizon, the derivative of the field $\partial _{+}\Psi
1096: _{p}^{(+)}\propto \left( x^{+}\right) ^{ip-1}$ diverges as $%
1097: x^{+}\rightarrow 0$, and this signals an infinity energy density, since the
1098: metric near the horizon has the regular form $ds^{2}\simeq -dx^{+}dx^{-}$. This
1099: fact was noted already in \cite{HS,Q1}. 
1100: 
1101: A natural way to cure the problem is to consider wave
1102: functions which are given by linear superpositions of the above basic
1103: solutions with different values of $p$. The problem is then to understand if
1104: general perturbations in the far past $t\ll -E^{-1}$ will evolve
1105: into the future and create an infinite energy density on the horizon, thus
1106: destabilizing the geometry. This problem is well known in the physics of
1107: black holes where, generically, Cauchy horizons are unstable to small
1108: perturbations of the geometry \cite{PenroseSimpson,CH}. We will show in the
1109: next section that the problem does not arise in this cosmological geometry.
1110: More precisely, we will show that perturbations which are localized in the $x
1111: $ direction in the far past $t\ll -E^{-1}$ (and which are therefore necessarily a
1112: superposition of our basic solutions (\ref{basicsol})) do not induce
1113: infinite energy densities on the horizons of the geometry, and can be
1114: continued smoothly into the other regions II$_{L,R}$ and I$_{out}$ of the
1115: geometry. 
1116: 
1117: This problem was also considered in \cite{Q2}, where a different conclusion
1118: was reached. We will discuss, at the end of section \ref{last}, the argument of 
1119: \cite{Q2} and describe why it does not apply to physically relevant bounded
1120: perturbations.
1121: 
1122: \subsection{Analysis of Infalling Matter}
1123: 
1124: For simplicity we will set, from now on, $E=1$ and $m=0$ and we will focus
1125: uniquely on the uncharged sector $n=0$, which corresponds to pure
1126: dimensional reduction. The three--dimensional action  
1127: for the massless scalar $\Psi$ reduces, in two dimensions, to 
1128: \begin{equation*}
1129: \int d^{2}x\sqrt{g}\Phi \left( \nabla \Psi \right) ^{2}\,.
1130: \end{equation*}
1131: The scalar field $\Psi $ is therefore conformally coupled, but has a non
1132: trivial coupling to the dilaton, and satisfies the equation of motion 
1133: \begin{equation}
1134: \square \Psi +\nabla \ln\Phi \cdot \nabla \Psi =0\,.  \label{EoM1}
1135: \end{equation}
1136: As in the previous section, 
1137: we analyze this equation in the background described by the vacuum
1138: solution (\ref{vacuumsol}), which is given in regions I$_{in,out}$ by 
1139: \begin{eqnarray}
1140: ds^{2} &=&-dt^{2}+\frac{t^{2}}{1+t^{2}}dx^{2}\,,  \label{vacuumSol} \\
1141: \Phi  &=&\sqrt{1+t^{2}}\,,  \notag
1142: \end{eqnarray}
1143: where $x^{\pm }=te^{\pm x}$. We have seen, in the previous section, that the
1144: solutions to (\ref{EoM1}) are known exactly in terms of Bessel functions,
1145: since (\ref{EoM1}) is equivalent to the equation 
1146: $\left( \square _{\mathrm{Flat}}-J^{2}\right) \Psi =0$,
1147: where $J=-i\frac{\partial }{\partial x}$ is the boost operator in the $%
1148: x^{\pm }$ plane and $\square _{\mathrm{Flat}}$ is the Laplacian with respect
1149: to the \textit{flat metric }$-dt^{2}+t^{2}dx^{2}\,$. A different and more
1150: general approach to the solution of (\ref{EoM1}), which is more in line with
1151: the literature on the stability of Cauchy horizons in black hole geometries,
1152: is to rewrite (\ref{EoM1}) in terms of the field 
1153: \begin{equation*}
1154: \Lambda =\Psi \sqrt{\Phi }\,.
1155: \end{equation*}
1156: A simple computation shows that equation (\ref{EoM1}) becomes 
1157: \begin{equation*}
1158: \square \Lambda +\frac{1}{4\Phi ^{2}}\left[ \left( \nabla \Phi \right)
1159: ^{2}-2\Phi \square \Phi \right] \Lambda =0\,.
1160: \end{equation*}
1161: We now use the fact that, in the vacuum solution (\ref{vacuumSol}), 
1162: equations (\ref{metric-eq}) and (\ref{Cfunct}) with $C=-1$ imply that
1163: $\square \Phi =-2\Phi ^{-3}$ and that $\left( \nabla \Phi
1164: \right) ^{2}=-1+\Phi ^{-2}$. Therefore
1165: the equation of motion becomes 
1166: \begin{equation}
1167: \square \Lambda +\frac{5-\Phi ^{2}}{4\Phi ^{4}}\Lambda =0\,.  \label{EoM2}
1168: \end{equation}
1169: 
1170: As customary, we write the above equation using conformal coordinates
1171: defined in region I$_{in}$. In the sequel, we follow closely the 
1172: beautiful work of Chandrasekar and Hartle \cite{CH}.
1173: We first define, in terms of the coordinates $%
1174: z^{\pm }$ in (\ref{confcoord}), the coordinates $u$, $v$ given by 
1175: \begin{eqnarray*}
1176: z^{+} &=&-e^{-v}\,, \\
1177: z^{-} &=&-e^{-u}\,.
1178: \end{eqnarray*}
1179: The coordinates $u$, $v$ are lightcone coordinates in the $s$, $x$ plane 
1180: \begin{eqnarray*}
1181: u &=&s+x\,, \\
1182: v &=&s-x\,,
1183: \end{eqnarray*}
1184: where $s$ is the usual \textit{tortoise} time coordinate. Recalling (\ref{confcoord}), $s$
1185: is given in terms of $t$ by the expression 
1186: \begin{equation}
1187: s=-\sqrt{1+t^{2}}+\frac{1}{2}\ln \frac{\sqrt{1+t^{2}}+1}{\sqrt{1+t^{2}}-1}\,
1188: \label{tortoise}
1189: \end{equation}
1190: and has asymptotic behaviour for large values of $\left| s\right| $ given by
1191: (see Figure \ref{fig9}) 
1192: \begin{eqnarray*}
1193: s &\sim &t\, ,\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
1194: \ \ \ \ \ \ \ \ \ \ \ \ \left( s\ll -1\right)  \\
1195: s &\sim &-\ln \left( -t\right) \, .\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
1196: \ \ \ \ \ \ \ \ \ \ \ \left( s\gg 1\right) 
1197: \end{eqnarray*}
1198: \begin{figure}
1199: \begin{center}
1200: \includegraphics{fig9.eps}
1201: \end{center}
1202: \caption{The \textit{tortoise} time coordinate $s$ versus $t$, together with the 
1203: asymptotic behaviours $t$ and $-\ln(-t)$ for $t\to -\infty$ and $t\sim 0$.
1204: \label{fig9}}
1205: \end{figure}
1206: The various coordinate systems are recalled in Figure \ref{fig11}a. The metric in
1207: region I$_{in}$ is given, from equation (\ref{confmetric}), by 
1208: $-dudv\,t^{2}\left( 1+t^{2}\right) ^{-1}$, and we conclude
1209: that equation (\ref{EoM2}) reads 
1210: \begin{equation*}
1211: \left( \frac{\partial ^{2}}{\partial s^{2}}-\frac{\partial ^{2}}{\partial
1212: x^{2}}\right) \Lambda \left( s,x\right) =V\left( s\right) \Lambda \left(
1213: s,x\right) \,,
1214: \end{equation*}
1215: where 
1216: \begin{equation}
1217: V=\frac{t^2}{4}\frac{\left(4-t^{2}\right)}{\left(1+t^{2}\right) ^{3}}\, .
1218: \label{potential}
1219: \end{equation}
1220: \begin{figure}
1221: \begin{center}
1222: \includegraphics{fig10.eps}
1223: \end{center}
1224: \caption{The scattering potential $V(s)$.\label{fig10}}
1225: \end{figure} 
1226: Finally, we Fourier transform the $x$ coordinate 
1227: \begin{equation*}
1228: \Lambda \left( s,x\right) =\int dp\,e^{ipx}\,\Lambda \left( s,p\right) 
1229: \end{equation*}
1230: to obtain the Schr\"{o}dinger--like equation 
1231: \begin{equation}
1232: \left( \frac{\partial ^{2}}{\partial s^{2}}+p^{2}\right) \Lambda \left(
1233: s,p\right) =V\left( s\right) \Lambda \left( s,p\right) \,.  \label{ODE}
1234: \end{equation}
1235: 
1236: Let us consider this differential equation in more detail. The potential $%
1237: V\left( s\right) $, shown in Figure \ref{fig10}, has the following asymptotic
1238: behaviour 
1239: \begin{eqnarray}
1240: V &\sim &e^{-2s}\,\,,\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
1241: \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \left( s\rightarrow \infty \right) 
1242: \label{asim} \\
1243: V &\sim &-\frac{1}{4s^{2}}\,\ ,\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
1244: \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \left( s\rightarrow -\infty \right)  
1245: \notag
1246: \end{eqnarray}
1247: and is therefore significant only in the region around $s\sim 0$. 
1248: The scattering potential connects the two regions with $s\ll -1$ and $s\gg 1$ (or $%
1249: t\ll -1$ and $t\sim 0$) where the field $\Lambda $ behaves
1250: essentially like a free scalar in two--dimensional flat space. The potential $%
1251: V$ then connects the past Minkowski region ($t\ll -1$ with metric $%
1252: -dt^{2}+dx^{2}$) to the future Milne wedge ($t\sim 0$ with metric $%
1253: -dt^{2}+t^{2}dx^{2}$).
1254: Since $V$ decays at infinity faster then $s^{-1}$, one can follow the usual
1255: theory of one--dimensional scattering. In particular, following \cite{CH}, we
1256: consider the solutions $F\left( s,p\right) $ and $P\left( s,p\right) $ of (%
1257: \ref{ODE}) which behave like pure exponentials $e^{-ips}$ respectively in
1258: the future $s\rightarrow \infty $ and in the past $s\rightarrow -\infty $%
1259: \begin{eqnarray}
1260: F\left( s,p\right)  &\sim &e^{-ips}\sim \left( -t\right) ^{ip}\,,\,\ \ \ \ \
1261: \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \left( s\rightarrow
1262: \infty \right)   \label{as} \\
1263: P\left( s,p\right)  &\sim &e^{-ips}\sim e^{-ipt}\,.\ \ \ \ \ \ \ \ \ \ \ \ \
1264: \ \ \ \ \ \ \ \ \ \ \ \ \ \ \left( s\rightarrow -\infty
1265: \right)   \notag
1266: \end{eqnarray}
1267: Therefore, when restoring the $x$ dependence $e^{ipx}$ of the full solution,
1268: we have, for $s\to\infty$, the plane waves
1269: \begin{eqnarray*}
1270: e^{ipx}F\left( s,p\right)  &\sim &e^{-ipv}\,,  \\
1271: e^{ipx}F\left( s,-p\right)  &\sim &e^{ipu}\,,
1272: \end{eqnarray*}
1273: and, for $s\to-\infty$,
1274: \begin{eqnarray*}
1275: e^{ipx}P\left( s,p\right)  &\sim &e^{-ipv}\,, \\
1276: e^{ipx}P\left( s,-p\right)  &\sim &e^{ipu}\,.
1277: \end{eqnarray*}
1278: As explained in the previous section, the problem is exactly soluble in terms of
1279: Bessel functions\footnote{%
1280: Recall that $J_{\nu }\left( z\right) =z^{\nu }\times \left( \mathrm{%
1281: entire\;function\;of\;}z\right) $. In this paper we choose \textit{%
1282: unconventionally} to take the branch cut of the logarithm along the negative
1283: imaginary axis, so that $z^{\nu }$ and $J_{\nu }\left( z\right) $ also have
1284: a cut on $\func{Re}z=0$, $\func{Im}z<0$. Moreover, the expressions like $%
1285: p^{ip}$ and $\sqrt{p}$ will always be defined using the same prescription
1286: for the logarithm.}, and one has the following explicit form of the
1287: functions $F$ and $P$%
1288: \begin{eqnarray*}
1289: F\left( s,p\right)  &=&\left( \frac{2}{p}\right) ^{ip}\,\Gamma \left(
1290: 1+ip\right) \,\,J_{ip}\left( -pt\right) \,\left( 1+t^{2}\right) ^{\frac{1}{4}%
1291: }\,, \\
1292: P\left( s,p\right)  &=&e^{\frac{i\pi }{4}-\frac{\pi p}{2}}\,\sqrt{\frac{\pi p%
1293: }{2}}\,H_{ip}^{\left( 1\right) }\left( -pt\right) \,\left( 1+t^{2}\right) ^{%
1294: \frac{1}{4}}\, ,
1295: \end{eqnarray*}
1296: where $t=t\left( s\right) $ is defined by (\ref{tortoise}). The function
1297: \begin{equation}
1298: H_{ip}^{\left( 1\right) }=\frac{1}{\sinh\left(
1299: \pi p\right)} \left( e^{\pi p}J_{ip}-J_{-ip}\right)
1300: \label{H1}\end{equation}
1301: is the Hankel function of the first kind, which has the correct asymptotic behaviour 
1302: to ensure that $P(s,p)$ has the desired properties.
1303: 
1304: It is useful,
1305: again following closely \cite{CH}, to consider the general analytic behaviour of $F$
1306: and $P$ as a function of the \textit{complex variable} $p$, at fixed time $s$%
1307: . First it is clear that we must have 
1308: \begin{eqnarray*}
1309: \overline{F\left( s,p\right) } &=&F\left( s,-\overline{p}\right) \,, \\
1310: \overline{P\left( s,p\right) } &=&P\left( s,-\overline{p}\right) \,.
1311: \end{eqnarray*}
1312: \begin{figure}
1313: \begin{center}
1314: \includegraphics{fig12.eps}
1315: \end{center}
1316: \caption{The analytic structure of $F(s,p)$ and $P(s,p)$ as functions of $p$.\label{fig12}}
1317: \end{figure}
1318: Moreover, we recall from \cite{CH} that, whenever the potential has asymptotic
1319: behaviour 
1320: \begin{eqnarray*}
1321: V &\sim &e^{-2\kappa _{-}s}\,,\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
1322: \ \ \ \ \ \ \ \ \ \ \ \ \left( s\rightarrow \infty \right)  \\
1323: V &\sim &e^{2\kappa _{+}s}\ ,\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
1324: \ \ \ \ \ \ \ \ \ \ \ \ \left( s\rightarrow -\infty \right) 
1325: \end{eqnarray*}
1326: then $F\left( s,p\right) $ is everywhere analytic, except for poles at $%
1327: p=i\kappa _{-}\,\mathbb{N}\ $(with $\mathbb{N}$ being the natural numbers $%
1328: 1,2,\cdots $), and $P\left( s,p\right) $ has poles at $p=-i\kappa _{+}%
1329: \mathbb{N}$. In our case $\kappa _{-}=1$ and, as expected, the function $F$
1330: has poles at $p=i\mathbb{N}$ due to the gamma function\footnote{%
1331: The function $\left( \frac{2}{p}\right) ^{ip}\,J_{ip}\left( -pt\right) $ is
1332: an entire function of $p$, as can be seen from the explicit power series
1333: representation of the Bessel function $J_{\nu }\left( z\right) $.} $\Gamma
1334: \left( 1+ip\right) $. On the other hand, for the case of the specific
1335: potential (\ref{potential}), one has $\kappa _{+}\rightarrow 0$, with the
1336: exponential behaviour replaced with a power law $s^{-2}$. This is a direct
1337: consequence of the fact that the metric distance from a point in region I$%
1338: _{in}$ to the horizon at $s\rightarrow \infty $ is finite, whereas it is
1339: infinite going to past infinity at $s\rightarrow -\infty $. Therefore, the
1340: poles at $-i\kappa _{+}\mathbb{N}$ become infinitely close and are
1341: effectively replaced by a branch cut along the negative imaginary $p$ axis,
1342: as one can check from the analytic expression for $P$. The features just
1343: described, which are shown in Figure \ref{fig12}, 
1344: are generic and do not depend on the details of the potential,
1345: but only on the asymptotic behaviour (\ref{asim}). 
1346: In particular, one has the same behaviour for fluctuations
1347: of \textit{any field}.
1348: 
1349: Let us now consider a general solution $\Lambda \left( s,p\right) $ of (\ref
1350: {ODE}). Since both $F\left( s,\pm p\right) $ and $P\left( s,\pm p\right) $
1351: are bases of the solutions of (\ref{ODE}), one must have that 
1352: \begin{eqnarray}
1353: \Lambda \left( s,p\right)  &=&V_{F}\left( p\right) F\left( s,p\right)
1354: +U_{F}\left( p\right) F\left( s,-p\right)   \notag \\
1355: &=&V_{P}\left( p\right) P\left( s,p\right) +U_{P}\left( p\right) P\left(
1356: s,-p\right) \,.  \label{past}
1357: \end{eqnarray}
1358: The coefficients $V_{F,P}$ and $U_{F,P}$ can be easily computed by recalling
1359: that, given two solutions $f$, $g$ of (\ref{ODE}), the Wronskian 
1360: \begin{equation*}
1361: \left[ f,g\right] =f\dot{g}-g\dot{f}
1362: \end{equation*}
1363: is independent of $s$, where the dot denotes differentiation with respect to 
1364: $s$.
1365: \begin{figure}
1366: \begin{center}
1367: \includegraphics{fig11.eps}
1368: \end{center}
1369: \caption{Figure (a) shows the various coordinate systems in region I$_{in}$.
1370: Figure (b) shows the various ways
1371: to define a solution $\Lambda$. One either gives the functions $V_P$, $U_P$ or the
1372: functions $V_F$, $U_F$, or, alternatively, the functions $\Lambda(s_0,x)$, 
1373: $\dot{\Lambda}(s_0,x)$ at some fixed time $s_0$ along the dotted line.\label{fig11}}
1374: \end{figure}
1375: In particular, one can easily show that 
1376: \begin{equation*}
1377: \left[ F\left( s,p\right) ,F\left( s,-p\right) \right] =\left[ P\left(
1378: s,p\right) ,P\left( s,-p\right) \right] =2ip\,.
1379: \end{equation*}
1380: Therefore we readily conclude that 
1381: \begin{eqnarray}
1382: V_{F}\left( p\right)  &=&\frac{1}{2ip}\left[ \Lambda \left( s,p\right)
1383: ,F\left( s,-p\right) \right] \,,  \label{future} \\
1384: U_{F}\left( p\right)  &=&-\frac{1}{2ip}\left[ \Lambda \left( s,p\right)
1385: ,F\left( s,p\right) \right] \,,  \notag
1386: \end{eqnarray}
1387: together with similar equations for $V_{P}$ and $U_{P}$.
1388: 
1389: \subsection{Behaviour at the Horizons\label{last}}
1390: 
1391: The meaning of the
1392: coefficients $V_{F}$ and $U_{F}$ is quite clearly understood by considering
1393: the value of the field $\Lambda $ on the horizons. Since we are in the limit 
1394: $s\rightarrow \infty $, we use the asymptotic form of $F\left( s,p\right) $
1395: to conclude that (see Figure \ref{fig11}b)
1396: \begin{equation}
1397: \Lambda \left( v,u\right) \rightarrow V_{F}\left( v\right) +U_{F}\left(
1398: u\right) \,,  \label{a1}
1399: \end{equation}
1400: where 
1401: \begin{equation*}
1402: V_{F}\left( v\right) =\int dp\,V_{F}\left( p\right) e^{-ipv}\,,\ \ \ \ \ \ \
1403: \ \ \ \ \ \ \ \ U_{F}\left( u\right) =\int dp\,U_{F}\left( p\right)
1404: e^{ipu}\,.
1405: \end{equation*}
1406: To compute the stress tensor due to the fluctuation in $\Lambda $, whose
1407: divergence usually signals the instability of the Cauchy horizon, we also
1408: must consider the derivatives of the field with respect to the coordinates $%
1409: z^{\pm }$ (the coordinates $u$, $v$ tend to $\infty $ on the horizons).
1410: Recalling that 
1411: \begin{equation*}
1412: \partial _{+}=-e^{v}\partial _{v}\, ,\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
1413: \ \ \ \ \ \ \ \ \ \ \ \ \ \partial _{-}=-e^{u}\partial _{u}\, ,
1414: \end{equation*}
1415: we conclude that 
1416: \begin{eqnarray}
1417: \partial _{+}\Lambda \left( v,u\right)  &\rightarrow &-e^{v}\partial
1418: _{v}V_{F}\left( v\right) \,,  \label{a2} \\
1419: \partial _{-}\Lambda \left( v,u\right)  &\rightarrow &-e^{u}\partial
1420: _{u}U_{F}\left( u\right) \,.  \notag
1421: \end{eqnarray}
1422: Note that both expressions (\ref{a1}) and (\ref{a2}) are valid \textit{only in
1423: the limit }$( u\rightarrow \infty \mathit{,\ }$ $v\mathit{\ }\mathrm{fixed}%
1424: ) $\textit{\ or }$\left( v\rightarrow \infty \mathit{,\ }u\mathit{\ }%
1425: \mathrm{fixed}\right) $\textit{, and are not approximate expressions for
1426: large but finite }$u$\textit{, }$v$\textit{. }This is because, although the
1427: asymptotic form (\ref{as}) is exactly valid in the limit $s\rightarrow
1428: \infty $, it becomes accurate for $s\gg \overline{s}\left( p\right) $, where
1429: the threshold value $\overline{s}\left( p\right) $ depends explicitly on $p$%
1430: . Given the explicit form of $F$, it is clear that $\overline{s}\left(
1431: p\right) \sim \ln p$ since only in this case is the argument of the Bessel
1432: function $-pt\ll 1$. Therefore, given a function $\Lambda $ which has
1433: support at all momenta, one is not uniformly in the asymptotic region for
1434: all momenta at any large but finite $s$. Only in the limit $s\rightarrow
1435: \infty $ we can use uniform convergence to deduce the expressions (\ref{a1})
1436: and (\ref{a2}), in the sense just described.
1437: 
1438: Finally, going back to the computation of the derivatives $\partial _{\pm }\Lambda $,
1439: one recalls from \cite{CH} that the problematic limits are given by $%
1440: \lim_{v\rightarrow \infty }\,\partial _{+}\Lambda $ and $\lim_{u\rightarrow
1441: \infty }\,\partial _{-}\Lambda $. Concentrating on the first one for
1442: simplicity, we see that we need to consider the expression 
1443: \begin{eqnarray}
1444: \lim_{v\rightarrow \infty }\,\partial _{+}\Lambda  &=&\lim_{v\rightarrow
1445: \infty }\,e^{v}\int dp\,V_{F}\left( p\right) \,ip\,e^{-ipv}  \label{problem}
1446: \\
1447: &=&\frac{1}{2}\lim_{v\rightarrow \infty }\,e^{v}\int dp\left[ \Lambda \left(
1448: s,p\right) ,F\left( s,-p\right) \right] \,e^{-ipv}\,.  \notag
1449: \end{eqnarray}
1450: The physical problem we need to address is the following. Let us
1451: assume that, at some time $s_{0}\ll 0$ in the past, much before the
1452: scattering potential $V\left( s\right) $ becomes relevant, our field $%
1453: \Lambda $ is in some specific configuration given by initial conditions on
1454: the value of $\Lambda $ and of its time derivative $\dot{\Lambda}$ 
1455: \begin{equation*}
1456: \Lambda \left( s_{0},x\right) \,,\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
1457: \ \dot{\Lambda}\left( s_{0},x\right) \,.
1458: \end{equation*}
1459: We will assume that the functions above are \textit{localized} as a function
1460: of $x$. More precisely, we shall demand that $\Lambda \left( s_{0},x\right) $
1461: and $\dot{\Lambda}\left( s_{0},x\right) $ are such that their evolution
1462: would not lead to an inconsistent behaviour on the horizon in the Milne wedge
1463: of \textit{flat space}, where $V=0$.
1464: It is important to realize that those are the general physical perturbation, since
1465: we are only demanding that they would be regular in flat space.
1466: Let us be more explicit. In the absence of the potential $V$, the solutions
1467: of the massless Klein--Gordon equations are just $\Lambda = \alpha(v) + \beta(u)$,
1468: and the initial conditions $\Lambda$ and $\dot\Lambda$ at
1469: $s_0$ are equivalent to giving the functions $\alpha$ and $\beta$ of the light--cone
1470: coordinates. Recall though that the coordinates which are regular across the horizon 
1471: are $z^\pm$, and therefore we must demand, in order to have a regular perturbation 
1472: in flat space, that $\alpha$ and $\beta$ be well--behaved \textit{as functions of}
1473: $z^\pm$. To see what this means in practice, consider a function of $z^+$ which
1474: has a nice power series expansion $a_0 + a_1 z^+ + \cdots$ around $z^+=0$.
1475: As a function of $v$ this reads  $\alpha(v) = a_0 - a_1 e^{-v} + \cdots$, and this shows
1476: that, aside from the constant part, the function $\alpha(v)$ must decay, for $v\to\infty$,
1477: at least as fast as $e^{-v}$. This fact imposes constraints on the Fourier transforms
1478: of $\alpha$ and $\beta$ and, in turn, on the functions $\Lambda(s_0,x)$ and $\dot{\Lambda}
1479: (s_0,x)$. In practice, it is sufficient to require that the Fourier transforms $\Lambda
1480: \left( s_{0},p\right) $ and $\dot{\Lambda} \left( s_{0},p\right) $ do not have
1481: poles in the strip $\left| \func{Im}p\right| <1$. Then the problematic limit 
1482: \begin{equation*}
1483: \frac{1}{2}\lim_{v\rightarrow \infty }\,e^{v}\int dp\left[ \Lambda \left(
1484: s_{0},p\right) \dot{F}\left( s_{0},-p\right) -\dot{\Lambda} \left( s_{0},p\right)
1485: F\left( s_{0},-p\right) \right] \,e^{-ipv}
1486: \end{equation*}
1487: can be computed by deforming the contour in the lower half of the complex
1488: plane, and is determined by the poles of $F\left( s,-p\right) $. In fact,
1489: the leading behaviour of the integral is controlled by the first pole at $-i$%
1490: , which gives an asymptotic behaviour for large $v\rightarrow \infty $ of $%
1491: e^{-v}$. The limit in the above expression then tends to a finite result.
1492: 
1493: In order to better understand the above result, let us review the standard
1494: argument for the instability of the Cauchy horizon, which was followed in \cite{Q2}.
1495: The problem is usually posed by giving initial conditions at past null infinity 
1496: (see also Figure \ref{fig11}b), by giving the functions 
1497: \begin{equation*}
1498: V_{P}\left( v\right) =\int dp\, V_{P}\left( p\right) e^{-ipv}\, ,\ \ \ \ \
1499: \ \ \ \ \ \ \ \ \ \ \ U_{P}\left( u\right) =\int dp\, U_{P}\left( p\right)
1500: e^{ipu}\, ,
1501: \end{equation*}
1502: on which we impose a regularity constraint (say, for simplicity, a
1503: localization condition like the one discussed above, now in the coordinates $%
1504: v$ and $u$). To determine $V_{F}$, $U_{F}$ in terms of $V_{P}$, $U_{P}$ one
1505: uses the expression (\ref{past}) for $\Lambda $ in the equation (\ref{future}%
1506: ) and obtains that 
1507: \begin{equation*}
1508: \left( 
1509: \begin{array}{c}
1510: V_{F}\left( p\right)  \\ 
1511: U_{F}\left( p\right) 
1512: \end{array}
1513: \right) =\left( 
1514: \begin{array}{cc}
1515: A\left( p\right)  & B\left( -p\right)  \\ 
1516: B\left( p\right)  & A\left( -p\right) 
1517: \end{array}
1518: \right) \left( 
1519: \begin{array}{c}
1520: V_{P}\left( p\right)  \\ 
1521: U_{P}\left( p\right) 
1522: \end{array}
1523: \right) \,,
1524: \end{equation*}
1525: where\footnote{%
1526: The coefficients $A\left( p\right) $ and $B\left( p\right) $ satisfy,
1527: generically, the conjugation relations $\overline{A\left( p\right) }=A\left( -\overline{p}\right)$
1528: and $\overline{B\left(p\right) }=B\left( -\overline{p}\right)$, 
1529: and the unitarity relation 
1530: $A\left( p\right) A\left( -p\right) -B\left( p\right) B\left( -p\right) =1$.
1531: Using (\ref{H1}), we have the explicit
1532: expression 
1533: \begin{eqnarray*}
1534: && A\left( p\right) = e^{\frac{\pi }{2}p+\frac{i}{4}\pi }\sqrt{\frac{\pi p}{2}}%
1535: \left( \frac{p}{2}\right) ^{ip}\frac{1}{\sinh \left( \pi p\right) }\frac{1}{%
1536: \Gamma \left( 1+ip\right) }\,, \\
1537: && B\left( p\right) = -e^{-\frac{\pi }{2}p+\frac{i}{4}\pi }\sqrt{\frac{\pi p}{2}}%
1538: \left( \frac{p}{2}\right) ^{-ip}\frac{1}{\sinh \left( \pi p\right) }\frac{1}{%
1539: \Gamma \left( 1-ip\right) }\,.
1540: \end{eqnarray*}
1541: Moreover, in general, the function $A(p)$ will have zeros on the \textit{positive
1542: imaginary $p$--axis} whenever the potential $V$ has a bound state. The explicit expressions
1543: above show that, for our particular potential, there are no bound states.
1544: } 
1545: \begin{eqnarray*}
1546: A\left( p\right)  &=&\frac{1}{2ip}\left[ P\left( s,p\right) ,F\left(
1547: s,-p\right) \right] \,, \\
1548: B\left( p\right)  &=&-\frac{1}{2ip}\left[ P\left( s,p\right) ,F\left(
1549: s,p\right) \right] \,.
1550: \end{eqnarray*}
1551: The above expression shows that both $A\left( p\right) $ and $B\left(
1552: p\right) $ have a cut along the negative imaginary $p$ axis (the limit of
1553: the poles at $-i\kappa _{+}n$ for $\kappa _{+}\rightarrow 0$). The 
1554: full analytic structure of $A(p)$ and $B(p)$ is shown in Figure \ref{fig13}. Since $%
1555: V_{F}\left( p\right) =V_{P}\left( p\right) A\left( p\right) +U_{P}\left(
1556: p\right) B\left( -p\right) $ we conclude that the leading pole of $V_P(p)A(p)$
1557: at $-i\kappa
1558: _{+}\rightarrow -i0$ determines the asymptotic behaviour of the integral in (%
1559: \ref{problem}) to be $e^{-\kappa _{+}v}$. Therefore, the limit $%
1560: \lim_{v\rightarrow \infty }e^{v\left( 1-\kappa _{+}\right) }$ diverges, thus
1561: naively signaling an instability of the horizon.
1562: \begin{figure}
1563: \begin{center}
1564: \includegraphics{fig13.eps}
1565: \end{center}
1566: \caption{The analytic structure of $A(p)$ and $B(p)$.\label{fig13}}
1567: \end{figure}
1568: 
1569: The above reasoning depends crucially on the assumption that the
1570: functions $U_{P}$, $V_{P}$ have a nice Fourier transform (analytic at least
1571: in the region $\left| \func{Im}p\right| <1$), so that the perturbation is
1572: localized on the null directions at past null infinity. If this is the case,
1573: though, expression (\ref{past}) shows that the function $\Lambda \left(
1574: s,x\right) $ is already delocalized (the Fourier transform has a branch cut
1575: due to $P\left( s,p\right) $) at \textit{any finite time }$s\ll -1$\textit{\
1576: much before the scattering potential}, and therefore will concentrate on the
1577: horizons and create an infinite stress tensor. This is clearly not the type
1578: of perturbation we want to focus on, which should be localized in space
1579: before they hit the potential $V$.
1580: Note that, at any finite time $s$, we cannot use, in equation (\ref{past}),
1581: the asymptotic form of $P(s,p)$ for all values of $p$. In fact, the 
1582: explicit form of the the function $P$ shows that one is in the 
1583: asymptotic regime for $|p|\gg s^{-1}$. Therefore, the crucial low
1584: momentum modes which determine the asymptotic behaviour of the wave
1585: function are \textit{never in the asymptotic region for any
1586: finite time $s$}. We then consider the requirement of
1587: localization of $\Lambda \left( s,x\right) $ at times $s\ll -1$ to be the
1588: correct and physically relevant boundary condition to study issues of
1589: stability of the geometry. The reasonings of this section then show that, in
1590: this sense, the Cauchy horizon in perfectly stable to small perturbations.
1591: 
1592: 
1593: \subsection{Discussion}
1594: 
1595: Let us conclude this section with some comments on future research.
1596: 
1597: We have shown that, in the orientifold cosmology, one can define particle--like
1598: perturbations which do not destabilize the Cauchy horizon, and therefore that
1599: the solution is stable against small variations of the background fields.
1600: All the work is done at the classical linearized level. The most pressing
1601: question for issues of stability is the study of the quantum stress--energy
1602: tensor for the field $\Psi$. This is a non--trivial problem due to the non--minimal
1603: coupling to the metric and the dilaton, and only some results are known in 
1604: the literature (see \cite{Noji, 2Dgravity} and references therein). On the other
1605: hand, the problem might be tractable since the exact wave--functions are known
1606: in this case.
1607: 
1608: 
1609: \section{Conclusion}
1610: 
1611: 
1612: In this paper we investigate the classical stability of a two--dimensional orientifold cosmology, 
1613: related to a time--dependent orbifold of flat space--time. We saw, with a specific 
1614: counter example, that the instability argument of Horowitz and Polchinski is not valid once the time--like
1615: orbifold singularity is interpreted as the boundary of space--time. In this specific example, we consider
1616: an exact shock--wave solution of two--dimensional dilaton gravity that uplifts to a distribution of matter 
1617: in the three--dimensional covering space. According to Horowitz and Polchinski this distribution should 
1618: interact gravitationally creating large black holes. Indeed, there are closed time--like curves in the covering space 
1619: geometry of the shock--wave, signaling the three--dimensional gravitational instability.
1620: However, such CTC's are not present if we interpret the singularity as a boundary of space--time, and accordingly 
1621: excise from the geometry the region behind it.
1622: 
1623: The other stability problem addressed in this work is related to cosmic censorship. The presence of naked 
1624: singularities with a Cauchy Horizon has lead many people to believe that such singularities 
1625: never form. Indeed, as for the Reissner--Nordstrom black hole one expects the Cauchy horizon to be unstable
1626: when crossed by matter. We analyze the propagation of a scalar field coupled to gravity in the geometry, 
1627: and show that any localized fluctuation at some finite time in the far past will not destabilize
1628: the horizon. The existence of the time--like singularity does {\it not} imply the break--down
1629: of predictability because of the conjectured duality between the singularity and orientifolds of 
1630: string theory. To leading approximation, the effect of the orientifolds is to enforce a boundary condition
1631: on the fields, which determines uniquely their evolution. 
1632: A more complete quantum understanding of this system is desirable.
1633: 
1634: 
1635: \section*{Acknowledgements}
1636: 
1637: We would like to thank J. Barbon, C. Herdeiro, F. Quevedo and I. Zavala 
1638: for useful discussions and correspondence. LC is supported by a Marie Curie 
1639: Fellowship under the European Commission's Improving Human Potential programme 
1640: (HPMF-CT-2002-02016). This work was partially supported by CERN under contract 
1641: CERN/FIS/43737/2001. 
1642: 
1643: \begin{thebibliography}{99}
1644: 
1645: \bibitem{Khoury}
1646: J.~Khoury, B.~A.~Ovrut, N.~Seiberg, P.~J.~Steinhardt and N.~Turok,
1647: \textit{From big crunch to big bang}, 
1648: \texttt{hep-th/0108187}.
1649: \\
1650: N.~Seiberg, \textit{From big crunch to big bang - is it possible?},
1651: \texttt{hep-th/0201039}.
1652: %%CITATION = HEP-TH 0108187;%%
1653: %%CITATION = HEP-TH 0201039;%%
1654: 
1655: \bibitem{VJ}
1656: V.~Balasubramanian, S.~F.~Hassan, E.~Keski-Vakkuri and A.~Naqvi,
1657: \textit{A space-time orbifold: A toy model for a cosmological singularity},
1658: \texttt{hep-th/0202187}.
1659: %%CITATION = HEP-TH 0202187;%%
1660: 
1661: \bibitem{CC}  L.~Cornalba and M.~S.~Costa, 
1662: \textit{A New Cosmological Scenario in String Theory}, 
1663: Phys.\ Rev.\ D \textbf{66} (2002) 066001, 
1664: \texttt{hep-th/0203031}. 
1665: %%CITATION = HEP-TH 0203031;%%
1666: 
1667: \bibitem{Nekrasov}
1668: N.~A.~Nekrasov,
1669: \textit{Milne universe, tachyons, and quantum group},
1670: \texttt{hep-th/0203112}.
1671: %%CITATION = HEP-TH 0203112;%%
1672: 
1673: \bibitem{Simon}
1674: J.~Simon,
1675: \textit{The geometry of null rotation identifications},
1676: \texttt{hep-th/0203201}.
1677: %%CITATION = HEP-TH 0203201;%%
1678: 
1679: \bibitem{Tolley:2002cv}
1680: A.~J.~Tolley and N.~Turok,
1681: \textit{Quantum fields in a big crunch / big bang spacetime},
1682: Phys.\ Rev.\ D {\bf 66} (2002) 106005,
1683: \texttt{hep-th/0204091}.
1684: %%CITATION = HEP-TH 0204091;%%
1685: 
1686: \bibitem{LMS1}  
1687: H.~Liu, G.~Moore and N.~Seiberg,
1688: \textit{Strings in a time-dependent orbifold},
1689: JHEP \textbf{0206} (2002) 045, 
1690: \texttt{hep-th/0204168}. 
1691: %%CITATION = HEP-TH 0204168;%%
1692: 
1693: \bibitem{Elitzur:2002rt}
1694: S.~Elitzur, A.~Giveon, D.~Kutasov and E.~Rabinovici,
1695: \textit{From big bang to big crunch and beyond},
1696: JHEP {\bf 0206} (2002) 017,
1697: \texttt{hep-th/0204189}.
1698: %%CITATION = HEP-TH 0204189;%%
1699: 
1700: \bibitem{CCK}  L.~Cornalba, M.~S.~Costa and C.~Kounnas, 
1701: \textit{A resolution of the cosmological singularity with orientifolds}, 
1702: Nucl.\ Phys.\ B \textbf{637} (2002) 378, 
1703: \texttt{hep-th/0204261}. 
1704: %%CITATION = HEP-TH 0204261;%%
1705: 
1706: \bibitem{Craps:2002ii}
1707: B.~Craps, D.~Kutasov and G.~Rajesh,
1708: \textit{String propagation in the presence of cosmological singularities},
1709: JHEP {\bf 0206} (2002) 053,
1710: \texttt{hep-th/0205101}.
1711: %%CITATION = HEP-TH 0205101;%%
1712: 
1713: \bibitem{Lawrence}  
1714: A.~Lawrence,
1715: \textit{On the instability of 3D null singularities}, 
1716: JHEP \textbf{0211} (2002) 019, 
1717: \texttt{hep-th/0205288}. 
1718: %%CITATION = HEP-TH 0205288;%%
1719: 
1720: \bibitem{Gordon:2002jw}
1721: C.~Gordon and N.~Turok,
1722: \textit{Cosmological perturbations through a general relativistic bounce},
1723: \texttt{hep-th/0206138}.
1724: %%CITATION = HEP-TH 0206138;%%
1725: 
1726: \bibitem{Martinec:2002xq}
1727: E.~J.~Martinec and W.~McElgin,
1728: \textit{Exciting AdS orbifolds},
1729: JHEP {\bf 0210} (2002) 050,
1730: \texttt{hep-th/0206175}.
1731: %%CITATION = HEP-TH 0206175;%%
1732: 
1733: \bibitem{LMS2}  
1734: H.~Liu, G.~Moore and N.~Seiberg, 
1735: \textit{Strings in time-dependent orbifolds}, 
1736: JHEP \textbf{0210} (2002) 031, 
1737: \texttt{hep-th/0206182}. 
1738: %%CITATION = HEP-TH 0206182;%%
1739: 
1740: \bibitem{Fabinger:2002kr}
1741: M.~Fabinger and J.~McGreevy,
1742: \textit{On smooth time-dependent orbifolds and null singularities},
1743: \texttt{hep-th/0206196}.
1744: %%CITATION = HEP-TH 0206196;%%
1745: 
1746: \bibitem{HP}  
1747: G.~T.~Horowitz and J.~Polchinski, 
1748: \textit{Instability of spacelike and null orbifold singularities},
1749: Phys.\ Rev.\ D \textbf{66} (2002) 103512, 
1750: \texttt{hep-th/0206228}. 
1751: %%CITATION = HEP-TH 0206228;%%
1752: 
1753: \bibitem{Buchel:2002kj}
1754: A.~Buchel, P.~Langfelder and J.~Walcher,
1755: \textit{On time-dependent backgrounds in supergravity and string theory},
1756: Phys.\ Rev.\ D {\bf 67} (2003) 024011,
1757: \texttt{hep-th/0207214}.
1758: %%CITATION = HEP-TH 0207214;%%
1759: 
1760: \bibitem{Hemming:2002kd}
1761: S.~Hemming, E.~Keski-Vakkuri and P.~Kraus,
1762: \textit{Strings in the extended BTZ spacetime},
1763: JHEP {\bf 0210} (2002) 006,
1764: \texttt{hep-th/0208003}.
1765: %%CITATION = HEP-TH 0208003;%%
1766: 
1767: \bibitem{Figueroa-O'Farrill:2002tb}
1768: J.~Figueroa-O'Farrill and J.~Simon,
1769: \textit{Supersymmetric Kaluza-Klein reductions of M2 and M5 branes},
1770: \texttt{hep-th/0208107}.
1771: %%CITATION = HEP-TH 0208107;%%
1772: 
1773: \bibitem{Hashimoto:2002nr}
1774: A.~Hashimoto and S.~Sethi,
1775: \textit{Holography and string dynamics in time-dependent backgrounds},
1776: Phys.\ Rev.\ Lett.\  {\bf 89} (2002) 261601,
1777: \texttt{hep-th/0208126}.
1778: %%CITATION = HEP-TH 0208126;%%
1779: 
1780: \bibitem{Simon:2002cf}
1781: J.~Simon,
1782: \textit{Null orbifolds in AdS, time dependence and holography},
1783: JHEP {\bf 0210} (2002) 036,
1784: \texttt{hep-th/0208165}.
1785: %%CITATION = HEP-TH 0208165;%%
1786: 
1787: \bibitem{Alishahiha:2002bk}
1788: M.~Alishahiha and S.~Parvizi,
1789: \textit{Branes in time-dependent backgrounds and AdS/CFT correspondence},
1790: JHEP {\bf 0210} (2002) 047,
1791: \texttt{hep-th/0208187}.
1792: %%CITATION = HEP-TH 0208187;%%
1793: 
1794: \bibitem{Dudas:2002dg}
1795: E.~Dudas, J.~Mourad and C.~Timirgaziu,
1796: \textit{Time and space dependent backgrounds from nonsupersymmetric strings},
1797: \texttt{hep-th/0209176}.
1798: %%CITATION = HEP-TH 0209176;%%
1799: 
1800: \bibitem{Satoh:2002nj}
1801: Y.~Satoh and J.~Troost,
1802: \textit{Massless BTZ black holes in minisuperspace},
1803: JHEP {\bf 0211} (2002) 042,
1804: \texttt{hep-th/0209195}.
1805: %%CITATION = HEP-TH 0209195;%%
1806: 
1807: \bibitem{Dolan:2002px}
1808: L.~Dolan and C.~R.~Nappi,
1809: \textit{Noncommutativity in a time-dependent background},
1810: Phys.\ Lett.\ B {\bf 551} (2003) 369,
1811: \texttt{hep-th/0210030}.
1812: %%CITATION = HEP-TH 0210030;%%
1813: 
1814: \bibitem{Cai:2002sv}
1815: R.~G.~Cai, J.~X.~Lu and N.~Ohta,
1816: \textit{NCOS and D-branes in time-dependent backgrounds},
1817: Phys.\ Lett.\ B {\bf 551} (2003) 178,
1818: \texttt{hep-th/0210206}.
1819: %%CITATION = HEP-TH 0210206;%%
1820: 
1821: \bibitem{Bachas:2002qt}
1822: C.~Bachas and C.~Hull,
1823: \textit{Null brane intersections},
1824: JHEP {\bf 0212} (2002) 035,
1825: \texttt{hep-th/0210269}.
1826: %%CITATION = HEP-TH 0210269;%%
1827: 
1828: \bibitem{Myers:2002bk}
1829: R.~C.~Myers and D.~J.~Winters,
1830: \textit{From D - anti-D pairs to branes in motion},
1831: JHEP {\bf 0212} (2002) 061,
1832: \texttt{hep-th/0211042}.
1833: %%CITATION = HEP-TH 0211042;%%
1834: 
1835: \bibitem{Okuyama:2002pc}
1836: K.~Okuyama,
1837: \textit{D-branes on the null-brane},
1838: \texttt{hep-th/0211218}.
1839: %%CITATION = HEP-TH 0211218;%%
1840: 
1841: \bibitem{Friedmann:2002gx}
1842: T.~Friedmann and H.~Verlinde,
1843: \textit{Schwinger meets Kaluza-Klein},
1844: \texttt{hep-th/0212163}.
1845: %%CITATION = HEP-TH 0212163;%%
1846: 
1847: \bibitem{BCKR}  
1848: M.~Berkooz, B.~Craps, D.~Kutasov and G.~Rajesh,
1849: \textit{Comments on cosmological singularities in string theory}, 
1850: \texttt{hep-th/0212215}. 
1851: %%CITATION = HEP-TH 0212215;%%
1852: 
1853: \bibitem{Fabinger:2002jn}
1854: M.~Fabinger and S.~Hellerman,
1855: \textit{Stringy resolutions of null singularities},
1856: \texttt{hep-th/0212223}.
1857: %%CITATION = HEP-TH 0212223;%%
1858: 
1859: \bibitem{Elitzur:2002vw}
1860: S.~Elitzur, A.~Giveon and E.~Rabinovici,
1861: \textit{Removing singularities},
1862: JHEP {\bf 0301} (2003) 017,
1863: \texttt{hep-th/0212242}.
1864: %%CITATION = HEP-TH 0212242;%%
1865: 
1866: \bibitem{HS}
1867: G.~T.~Horowitz and A.~R.~Steif,
1868: \textit{Singular String Solutions With Nonsingular Initial Data},
1869: Phys.\ Lett.\ B {\bf 258} (1991) 91.
1870: %%CITATION = PHLTA,B258,91;%%
1871: 
1872: \bibitem{SimonFigueroa}
1873: J.~Figueroa-O'Farrill and J.~Simon,
1874: \textit{Generalized supersymmetric fluxbranes},
1875: JHEP {\bf 0112} (2001) 011,
1876: \texttt{hep-th/0110170}.
1877: %%CITATION = HEP-TH 0110170;%%
1878: 
1879: \bibitem{PenroseSimpson}  
1880: M.~Simpson and R.~Penrose,
1881: \textit{\em Internal Instability in a Reissner-Nordstrom Black Hole},
1882: Int.\ Journ.\ Theor.\ Physics \textbf{7} (1973) 183.
1883: 
1884: \bibitem{CH} 
1885: S.~Chandrasekhar and J.B.~Hartle, 
1886: \textit{\em On Crossing the Cauchy Horizon of a Reissner--Nordstrom Black--Hole},
1887:  Proc.\ R.\ Soc.\ Lond.\ \textbf{A 384} (1982) 301-315.
1888: 
1889: \bibitem{KounnasLust}  
1890: C.~Kounnas and D.~Lust, 
1891: \textit{Cosmological string backgrounds from gauged WZW models}, 
1892: Phys.\ Lett.\ B \textbf{289} (1992) 56, 
1893: \texttt{hep-th/9205046}. 
1894: %%CITATION = HEP-TH 9205046;%%
1895: 
1896: \bibitem{Q0}
1897: C.~Grojean, F.~Quevedo, G.~Tasinato and I.~Zavala,
1898: \textit{Branes on charged dilatonic backgrounds: Self-tuning, Lorentz  violations and cosmology},
1899: JHEP {\bf 0108} (2001) 005,
1900: \texttt{hep-th/0106120}.
1901: %%CITATION = HEP-TH 0106120;%%
1902: 
1903: \bibitem{Q1}
1904: C.~P.~Burgess, F.~Quevedo, S.~J.~Rey, G.~Tasinato and I.~Zavala,
1905: \textit{Cosmological spacetimes from negative tension brane backgrounds},
1906: JHEP {\bf 0210} (2002) 028, 
1907: \texttt{hep-th/0207104}.
1908: %%CITATION = HEP-TH 0207104;%%
1909: 
1910: \bibitem{Matschull}  
1911: H.~J.~Matschull,
1912: \textit{Black hole creation in 2+1-dimensions},
1913: Class.\ Quant.\ Grav.\ \textbf{16} (1999) 1069, 
1914: \texttt{gr-qc/9809087}. 
1915: %%CITATION = GR-QC 9809087;%%
1916: 
1917: \bibitem{CGHS}  
1918: C.~G.~Callan, S.~B.~Giddings, J.~A.~Harvey and A.~Strominger,
1919: \textit{Evanescent Black Holes},
1920: Phys.\ Rev.\ D \textbf{45} (1992) 1005, 
1921: \texttt{hep-th/9111056}. 
1922: %%CITATION = HEP-TH 9111056;%%
1923: 
1924: \bibitem{Gott}  
1925: J.~R.~Gott, 
1926: \textit{Closed Timelike Curves Produced By Pairs Of Moving Cosmic Strings: Exact Solutions}, 
1927: Phys.\ Rev.\ Lett.\ \textbf{66} (1991) 1126. 
1928: %%CITATION = PRLTA,66,1126;%%
1929: 
1930: \bibitem{Deser}  
1931: S.~Deser, R.~Jackiw and G.~'t Hooft,
1932: \textit{Physical cosmic strings do not generate closed timelike curves},
1933: Phys.\ Rev.\ Lett.\ \textbf{68} (1992) 267. 
1934: %%CITATION = PRLTA,68,267;%%
1935: 
1936: \bibitem{Kabat}  
1937: D.~N.~Kabat,
1938: \textit{Conditions For The Existence Of Closed Timelike Curves In (2+1) Gravity},
1939: Phys.\ Rev.\ D \textbf{46} (1992) 2720. 
1940: %%CITATION = PHRVA,D46,2720;%%
1941: 
1942: \bibitem{Gott2}  
1943: M.~P.~Headrick and J.~R.~Gott, 
1944: \textit{(2+1)-Dimensional Space-Times Containing Closed Timelike Curves},
1945: Phys.\ Rev.\ D \textbf{50} (1994) 7244. 
1946: %%CITATION = PHRVA,D50,7244;%%
1947: 
1948: \bibitem{BTZ}  
1949: M.~Banados, C.~Teitelboim and J.~Zanelli, 
1950: \textit{The Black Hole In Three-Dimensional Space-Time}, 
1951: Phys.\ Rev.\ Lett.\ \textbf{69} (1992) 1849, 
1952: \texttt{hep-th/9204099}. 
1953: %%CITATION = HEP-TH 9204099;%%
1954: 
1955: 
1956: \bibitem{ADSagnotti}  
1957: I.~Antoniadis, E.~Dudas and A.~Sagnotti,
1958: \textit{Supersymmetry breaking, open strings and M-theory},
1959: Nucl.\ Phys.\ B  \textbf{544} (1999) 469, 
1960: \texttt{hep-th/9807011}. 
1961: %%CITATION = HEP-TH 9807011;%%
1962: 
1963: \bibitem{Kachru}  
1964: S.~Kachru, J.~Kumar and E.~Silverstein,
1965: \textit{Orientifolds, RG flows, and closed string tachyons},
1966: Class.\ Quant.\ Grav.\ \textbf{17} (2000) 1139, 
1967: \texttt{hep-th/9907038}. 
1968: %%CITATION = HEP-TH 9907038;%%
1969: 
1970: \bibitem{Noji}
1971: S.~Nojiri and S.~D.~Odintsov,
1972: \textit{Quantum dilatonic gravity in d = 2, 4 and 5 dimensions},
1973: Int.\ J.\ Mod.\ Phys.\ A \textbf{16}, 1015 (2001),
1974: \texttt{hep-th/0009202}.
1975: %%CITATION = HEP-TH 0009202;%%
1976: 
1977: \bibitem{2Dgravity}  
1978: D.~Grumiller, W.~Kummer and D.~V.~Vassilevich,
1979: \textit{Dilaton gravity in two dimensions}, 
1980: Phys.\ Rept.\ \textbf{369} (2002) 327, 
1981: \texttt{hep-th/0204253}. 
1982: %%CITATION = HEP-TH 0204253;%%
1983: 
1984: \bibitem{Kunstatter}  
1985: D.~Louis-Martinez and G.~Kunstatter,
1986: \textit{On Birckhoff's theorem in 2-D dilaton gravity}, 
1987: Phys.\ Rev.\ D \textbf{49} (1994) 5227. 
1988: %%CITATION = PHRVA,D49,5227;%%
1989: 
1990: \bibitem{Steif}  
1991: A.~R.~Steif,
1992: \textit{The Quantum Stress Tensor In The Three-Dimensional Black Hole},
1993: Phys.\ Rev.\ D \textbf{49} (1994) 585, 
1994: \texttt{gr-qc/9308032}. 
1995: %%CITATION = GR-QC 9308032;%%
1996: 
1997: \bibitem{Q2}
1998: C.~P.~Burgess, P.~Martineau, F.~Quevedo, G.~Tasinato and I.~Zavala C.,
1999: \textit{Instabilities and particle production in S-brane geometries}, 
2000: \texttt{hep-th/0301122}.
2001: %%CITATION = HEP-TH 0301122;%%
2002: 
2003: \end{thebibliography}
2004: 
2005: 
2006: \end{document}
2007: 
2008: