1: \documentclass[letterpaper]{JHEP3}
2: %\documentstyle[preprint,eqsecnum,aps,prd,]{revtex}
3: %\draftcurrent
4: %\input{epsf}
5: \usepackage{epsfig}
6: \usepackage{graphicx}
7: %\tighten
8: \newcommand{\sect}[1]{\section{#1}\setcounter{equation}{0}}
9: \def\theequation{\thesection.\arabic{equation}}
10: \def\thesection
11: {\arabic{section}}
12: \newcommand{\labell}[1]{{}_{#1}\qquad\label{#1}} %\label{#1}}
13: \newcommand{\reef}[1]{(\ref{#1})}
14: \newcommand{\ie}{{\it i.e.,}\ }
15: \newcommand{\eg}{{\it e.g.,}\ }
16: \newcommand{\ssc}{\scriptscriptstyle}
17: \newcommand{\GN}{G_{\ssc N}}
18: \newcommand{\rr}{\rho}
19: \def\IR{{\hbox{{\rm I}\kern-.2em\hbox{\rm R}}}}
20: \newcommand{\be}{\begin{equation}}
21: \newcommand{\ee}{\end{equation}}
22: \newcommand{\beq}{\begin{equation}}
23: \newcommand{\eeq}{\end{equation}}
24: \newcommand{\beqa}{\begin{eqnarray}}
25: \newcommand{\eeqa}{\end{eqnarray}}
26: \newcommand{\prt}{\partial}
27: \newcommand{\Lam}{\Lambda}
28: \newcommand{\lc}{\ell}
29: \newcommand{\tlc}{\tilde{\lc}}
30:
31: %====================================================================+
32: \preprint{{\small hep-th/0303035}} \keywords{S-branes, unstable
33: D-branes, rolling tachyons.}
34:
35: \title{SD-brane gravity fields and rolling tachyons}
36:
37: \author{ Fr\'ed\'eric Leblond\footnote{E-mail: {\tt
38: fleblond@perimeterinstitute.ca}} \ and Amanda W.
39: Peet\footnote{E-mail: {\tt peet@physics.utoronto.ca}}\\ $^{*}$
40: Department of Physics, McGill University, Montr\'eal, Qu\' ebec
41: H3A 2T8 Canada
42: \\ $^{*}$ Perimeter Institute for
43: Theoretical Physics, Waterloo, Ontario N2J 2W9 Canada \\
44: $^{\dagger}$ Radcliffe Institute, Harvard University, Cambridge,
45: MA 02138 USA \\ $^{\dagger}$ Department of Physics,
46: University of Toronto, Toronto, Ontario M5S 1A7 Canada \\
47: $^{\dagger}$ Cosmology and Gravity Program, Canadian Institute for
48: Advanced Research}
49:
50: \date{March, 2003}
51: \abstract{S(pacelike)D-branes are objects arising naturally in
52: string theory when Dirichlet boundary conditions are imposed on
53: the time direction. SD-brane physics is inherently
54: time-dependent. Previous investigations of gravity fields of
55: SD-branes have yielded undesirable naked spacelike singularities.
56: We set up the problem of coupling the most relevant open-string
57: tachyonic mode to massless closed-string modes in the bulk, with
58: backreaction and Ramond-Ramond fields included. We find solutions
59: numerically in a self-consistent approximation; our solutions are
60: naturally asymptotically flat and time-reversal asymmetric. We
61: find completely nonsingular evolution; in particular, the dilaton
62: and curvature are well-behaved for all time. The essential
63: mechanism for spacetime singularity resolution is the inclusion of
64: full backreaction between the bulk fields and the rolling tachyon.
65: Our analysis is not the final word on the story, because we have
66: to make some significant approximations, most notably homogeneity
67: of the tachyon on the unstable branes. Nonetheless, we provide
68: significant progress in plugging a gaping hole in prior
69: understanding of the gravity fields of SD-branes.}
70:
71: \keywords{S-branes, rolling tachyons, singularity resolution}
72:
73: %====================================================================+
74: \begin{document}
75: \setcounter{footnote}{0}
76: %====================================================================+
77: \section{Introduction}\label{section:introduction}
78:
79: SD-branes in string theory were first studied by Gutperle and
80: Strominger in ref.~\cite{gutperle}. They were introduced as
81: objects arising when Dirichlet boundary conditions on open strings
82: are put on the time coordinate, as well as on spatial coordinates.
83: SD-branes are not supersymmetric objects, which makes them hard to
84: handle but potentially very interesting. The boundary conditions
85: for SD-branes imply that they live for only an ``instant'' of
86: time, and so the worldvolume is purely spatial. SD-branes should
87: not be confused with instantons, because they live out their lives
88: in Lorentzian signature context. Recent discussion of the relation
89: between SD-branes and instantons may be found in section 6 of
90: ref.~\cite{andylastweek}.
91:
92: SD-branes are especially interesting objects to study in the
93: context of tachyon condensation, which will be the arena of our
94: investigation. SD-branes are indeed inherently related to the
95: general study of time dependence in string theory. One of the
96: original goals of ref.~\cite{gutperle} was in fact to seek
97: examples of gauge/gravity dualities where a time direction on the
98: gravity side is holographically reconstructed by a Euclidean field
99: theory.\footnote{An attempt to find a realization for such a
100: duality is the dS/CFT correspondence \cite{dscft} -- see also
101: ref.~\cite{tale} for an extensive list of references.}
102:
103: There have been several investigations of SD-branes since they
104: were introduced \cite{gutperle2, KMP, strominger, buchel, roy,
105: sbranes, andylastweek}. Most of them involve taking the limit
106: $g_s\rightarrow 0$, the regime in which perturbative string
107: computations can be done. As the tachyon rolls down its potential
108: hill, there is a divergence in production of higher mass open
109: string modes \cite{strominger,strominger3,andylastweek}. This
110: divergence occurs before the tachyon gets to the bottom of the
111: potential well, as it must because there are no perturbative open
112: string excitations around the true minimum of the tachyon
113: potential \cite{vacuum}. Also, the time taken to convert the
114: energy of the rolling tachyon into these massive open string modes
115: is of order\footnote{In later sections we will see that our
116: calculation, which includes gravity backreaction, agrees with this
117: in the sense that the time it takes for the tachyon to decay only
118: slightly depends on the particular value of $g_{s}$.}
119: ${\cal{O}}(g_{s}^{0})$ . This analysis was done for a single
120: SD-brane using CFT methods; analysis for production of massive
121: closed string modes was also done \cite{OS,maldacena}.
122:
123: One aspect of SD-brane physics has become clear: that the
124: decoupling limit applied to SD-branes is not a smooth limit like
125: it is for regular D-branes. In particular, as $g_s\rightarrow 0$
126: the brane tachyon becomes decoupled from bulk modes, which were
127: however the most natural modes into which the initial energy of
128: the unstable brane should decay. Then, the endpoint of the
129: rolling tachyon must include a somewhat mysterious substance
130: called ``tachyon matter'' \cite{sen2}. Consideration of the full
131: problem with $g_s$ finite would presumably eliminate the need for
132: mysterious tachyon matter; this was in fact part of our motivation
133: for this work. Regarding production of closed string massive
134: modes, at very small $g_s$, it seemed that there was some debate
135: \cite{OS,maldacena} about the form of a divergence.
136: %
137: The results from ref.~\cite{maldacena} make clear that the
138: divergence depends on the number of dimensions transverse to the
139: decaying unstable brane: for unstable D$p$-branes with $p<2$ there
140: was a need to invoke a cutoff to get a finite result.
141: %
142: In any case, unstable brane decay should presumably be a physically
143: smooth process for $\{g_s,\ell_s\}$ finite.
144:
145: The point of view that we will be taking is to consider a system of
146: $N$ SD-branes, with $g_sN$ large, study the overall centre-of-mass
147: tachyon, and couple it to bulk massless closed string modes.
148: Obviously, it would be nice to understand the full problem including
149: coupling to all massive open and closed string modes, but this is a
150: hard problem beyond our reach. We will make a beginning here with a
151: quantitative analysis involving only the lowest modes in each of the
152: open and closed string sectors. We believe that our approach, while
153: ``lowbrow'' by comparison to SFT computations, already shows some very
154: interesting physics.
155:
156: The punchline of our paper will be this: we find nonsingular solutions
157: for the evolution of the open string tachyon coupled to bulk
158: supergravity modes. This plugs a gaping hole in our previous
159: understanding \cite{gutperle,gutperle2,KMP} of the supergravity fields
160: arising from a large number of SD-branes. All previous attempts at
161: describing SD-branes in the context of supergravity had found that the
162: corresponding solutions were plagued with naked spacelike
163: singularities. We find that resolution of these singularities is
164: achieved in a conceptually simple way: by including full backreaction
165: on the rolling tachyon.
166:
167: Our investigation can also be considered to shed light on the
168: question of tachyon cosmology \cite{gibbons} including
169: Ramond-Ramond fields, the effect of which was ignored in previous
170: investigations. Tachyon cosmology itself may not yet provide
171: realistic models for inflation, nonetheless, see recent work
172: including, {\it e.g.}, refs.~\cite{cosmotachyonK,cosmotachyon}.
173: One of the reasons is that, in the low-energy actions used to
174: describe tachyon cosmology dynamics, there is only one length
175: scale --- the string length. It would certainly be interesting if
176: a mechanism generating a lower scale for inflation were found
177: within this context. Also, the behavior of inhomogeneities during
178: the later stage of the roll of the tachyon may be a general
179: problem \cite{cosmotachyonK}.
180: %
181: Tachyon cosmology involving brane-antibrane annihilation may be
182: relevant only to a pre-inflationary period, but it is interesting to
183: analyze the dynamics from the ``top-down'' perspective in string
184: theory.
185:
186: The plan of our paper is as follows. We begin by reviewing in
187: section \ref{section:review} the previous work on gravity fields
188: of SD-branes, and commenting on the nature of singularities found
189: in the past. In section \ref{subsection:ssss} we discuss
190: pathologies of the solutions found in refs.~\cite{gutperle2,KMP},
191: where only supergravity fields were considered. In section
192: \ref{subsection:buchel} we move to discussing ref.~\cite{buchel},
193: in which an unstable brane probe was coupled to a $d=4$ SD0-brane
194: supergravity background; we generalized their arguments but still
195: find generic singularities in the probe approximation (exceptions
196: are considered in Appendix~\ref{regular}).
197: %
198: In section \ref{section:newsetup} we set up the full backreacted
199: problem of interest. The equations are naturally highly nonlinear,
200: and since backreaction is essential we have no desire to ignore it or
201: treat it perturbatively. We need to use a particular homogeneous
202: ansatz to facilitate solution of the equations of motion, and we
203: discuss implications of the ansatz.
204: %
205: In section \ref{section:numeric} we demonstrate numerical solutions of
206: our backreaction-inclusive equations, and interpret qualitative
207: features found. In particular, we follow carefully the evolution of
208: both the dilaton and curvature invariants.
209: %
210: Lastly, in the Discussion section we summarize our conclusions, open
211: issues, and directions of future work.
212:
213: %====================================================================+
214: \section{Singular supergravity SD-branes: Review}\label{section:review}
215:
216: In the paper \cite{gutperle}, a small number of supergravity
217: solutions, thought to be appropriate for a large number $N$ of
218: SD-branes, were presented. Subsequently, the class of supergravity
219: solutions was widened considerably, simultaneously by two groups
220: (refs.~\cite{gutperle2} and \cite{KMP}). A later paper showed that
221: these two sets of solutions were equivalent \cite{roy}, by matching
222: boundary conditions on asymptotic fields and finding the coordinate
223: transformation explicitly.
224:
225: %--------------------------------------------------------------------+
226: \subsection{Sourceless SD-brane supergravity
227: solutions}\label{subsection:ssss}
228:
229: The convention of ref.~\cite{gutperle}, which we will use, is that
230: SD$p$-branes have $(p{+}1)$ worldvolume coordinates. We call
231: these ${\vec{y}}$. There is also the time coordinate $t$, and the
232: $(8{-}p)$ overall transverse coordinates ${\vec{x}}$.
233:
234: The most general SD$p$-brane supergravity solutions of
235: refs.~\cite{gutperle2} and \cite{KMP} can then be written in
236: string frame as follows,
237: \begin{eqnarray}
238: \label{bigmetric} ds^{2}_{{\scriptscriptstyle
239: Sp}}&=&F(t)^{1/2}\beta(t)^G\alpha(t)^H
240: \left(-dt^2+t^2 dH^{2}_{8-p}({\vec{x}})\right) \nonumber\\
241: %
242: && +F(t)^{-1/2}\left[\sum_{i=2}^{p+1}
243: \left({\beta(t)\over\alpha(t)}\right)^{-k_i}(dy^i)^2
244: +\left({\beta(t)\over\alpha(t)}\right)^{k_{1}+{\tilde{k}}}
245: (dy^{1})^2\right]\, , \nonumber\\
246: %
247: e^{2\Phi} &=& F(t)^{3-p\over2}
248: \left({\beta(t)\over\alpha(t)}\right)^{-\sum_{i=2}^{p+1}k_i}\ ,
249: \label{solupb}\\
250: %
251: C^{(p+1)}&=& \sin\theta\cos\theta\,{C(t)\over F(t)}\, dy^{1}\wedge
252: \cdots\wedge dy^{p+1}\ ,\nonumber
253: \end{eqnarray}
254: %%
255: where
256: %%
257: \begin{equation}
258: C(t) \equiv \left({\beta(t)\over\alpha(t)}\right)^{k_1}
259: -\left({\beta(t)\over\alpha(t)}\right)^{{\tilde{k}}}\,,\qquad
260: F(t)\equiv
261: \cos^2\theta\left({\beta(t)\over\alpha(t)}\right)^{{\tilde{k}}}
262: +\sin^2\theta\left({\beta(t)\over\alpha(t)}\right)^{k_1} \ ,
263: \label{useful}
264: \end{equation}
265: %%
266: and
267: %
268: \begin{equation}\label{alphabeta}
269: \alpha(t)\equiv 1+\left({\frac{\omega}{t}} \right)^{7-p}\,,\quad
270: \beta(t)\equiv 1-\left({\frac{\omega}{t}} \right)^{7-p}\,.
271: \end{equation}
272: %
273: The supergravity equations of motion will be satisfied when the
274: exponents satisfy the constraints,
275: %%
276: \begin{eqnarray}
277: {\tilde{k}}^2+\sum_{i=1}^{p+1}k_i^2+{7-p\over4}(H-G)^2-4
278: {8-p\over 7-p}&=&0\ , \nonumber\\
279: {\tilde{k}}+\sum_{i=1}^{p+1}k_i-{7-p\over2}(H-G)&=&0\ ,
280: \label{solve}\\
281: H+G-{4\over 7-p}&=&0\ . \nonumber
282: \end{eqnarray}
283: %%
284:
285: The general metric above is not isotropic in the worldvolume
286: directions ${\vec{y}}$. However, from a microscopic point of
287: view, one expects that the supergravity solution would have an
288: isotropic worldvolume. Isotropy in the worldvolume will be
289: restored in the above solution for the choice
290: $-k_2=\cdots=-k_{p+1}=k_{1}+{\tilde{k}}\equiv n$. Unfortunately,
291: the isotropy requirement excludes SD-brane solutions with regular
292: Cauchy horizon. In fact, the curvature invariants associated to
293: all isotropic solutions diverge at $t=\pm \omega$ and $t=0$.
294: Consequently, although they possess the right symmetries and
295: ``charge'', these solutions do not appear to be well-defined.
296:
297: Besides, SD-branes should be represented by solutions with,
298: roughly, three distinct regions: the infinite past with incoming
299: radiation only in the form of massless closed strings, an
300: intermediate region with both open and closed strings, and finally
301: the infinite future with dissipating outgoing radiation in the
302: form of massless closed strings. But the supergravity solutions of
303: refs.~\cite{gutperle,gutperle2,KMP} cannot represent this process,
304: because there are no rules for deciding how to go through the
305: singular regions (see, however, Appendix~\ref{regular} where we
306: show that some anisotropic solutions avoid the pathology).
307:
308: Nevertheless, the isotropic solutions have a positive feature
309: worth noting: they have the correct asymptotics at large time. In
310: the limit that the functions $\alpha(t)$ and $\beta(t)$ become
311: trivial, part of the metric is simply the Milne universe: flat
312: Minkowski space foliated by hyperbolic sections, \beq
313: \lim_{t\rightarrow \pm \infty}ds^{2}_{{\scriptscriptstyle Sp}}=
314: -dt^2+t^2 dH^2_{8-p}({\vec x}) + \sum_{i=1}^{p+2} (d{y}^i)^2 .
315: \eeq On the other hand, there are (at least) two reasons to
316: suspect that the above solutions are not the final word in the
317: SD-brane supergravity story. The first is that there are one too
318: many parameters in the solution, by comparison to expectations
319: from microscopics of SD-branes \cite{KMP}. A possible explanation
320: for this may be as follows. In the rest of our paper, we will be
321: showing that the full coupling/backreaction between the
322: open-string tachyon and the closed-string bulk modes is crucial
323: for resolution of spacetime singularities. It is possible that
324: the freedom in the supergravity solutions may correspond to a
325: freedom in picking boundary conditions for the rolling tachyon ---
326: the coupling to which was not included in
327: refs.~\cite{gutperle,gutperle2,KMP}.
328:
329: The most noticeable negative feature of the above solutions is
330: that the isotropic solutions are nakedly singular. Quite
331: generally, nakedly singular spacetimes arising in low-energy
332: string theory come under immediate suspicion, even though they are
333: solutions to the supergravity field equations. No-hair theorems
334: are usually what we rely on in order to be sure that we have the
335: unique supergravity solution, but no-hair theorems are never valid
336: for solutions with {\em naked} spacetime singularities. It is
337: worth noting that it has been shown with an explicit
338: counterexample \cite{emparan} that even no-hair theorems
339: themselves fail for black holes in $d=5$ with mass and angular
340: momentum --- and hence the idea of no-hair theorems in all higher
341: dimensions is under suspicion. (Nonetheless, with particular
342: assumptions about field couplings, no-hair theorems can be proven
343: for static asymptotically flat dilaton black holes
344: \cite{gibbonshair}. Also, uniqueness of the {\em supersymmetric}
345: rotating BMPV \cite{BMPV} black hole in $d=5$ has been proven
346: \cite{reallbmpv}.) Even if a no-hair theorem appropriate to the
347: supergravity theory involving SD-branes could be proven, however,
348: the above solutions we have reviewed would be ruled out as
349: candidates because their singularities are uncontrollably nasty.
350: So we have to look elsewhere.
351:
352: Let us take a brief sidetrip here to comment on the singularity
353: story for supergravity fields of $N{\gg}1$ regular D-branes with
354: timelike worldvolume. Certainly, the geometry of BPS D3-branes is
355: nonsingular, and there are several other pretty situations known
356: in the literature where branes ``melt'' into fluxes: the sources
357: are no longer needed. However, the disappearance of D-brane
358: sources for supergravity fields only occurs when the branes are
359: BPS. If any energy density above BPS is added to these systems,
360: singularities reappear: this certainly happens for the D3-brane
361: system. Also, in the low-energy approximation to string theory,
362: it is misleading to think of supergravity fields of D$p$-branes as
363: simple condensates of massless closed string modes. The reason is
364: directly analogous to the fact that the Coulomb field of an
365: electron cannot be a photon condensate because the photon is
366: transverse. Similarly, Coulomb Ramond-Ramond fields of
367: D$p$-branes cannot be represented by supergravity fields
368: alone.\footnote{This is the case even though D$p$-branes are
369: ``solitonic'' in string theory while electrons are fundamental in
370: QED. The straightforward argument we use here depends only on the
371: couplings of the charge-carrying objects to the bulk gauge fields.
372: We thank Abhay Ashtekar for a discussion on this issue.} This
373: river runs deeper: in the decoupling limit, resolution of dilaton
374: and curvature singularities for $p\not =3$ D$p$-branes is in fact
375: {\em provided} by the gauge theory on the D-branes \cite{IMSY,
376: awpconfproc}.
377:
378: Let us now get back to our SD-branes. The supergravity situation
379: looks similar to that for non-BPS (ordinary) D-branes: it seems that
380: brane modes will be required for singularity resolution. Therefore,
381: we are motivated to try to solve all problems with prior candidate
382: SD-brane spacetimes by solving the coupled system of brane tachyon
383: plus bulk supergravity fields with full backreaction.
384:
385: %--------------------------------------------------------------------+
386: \subsection{Unstable brane probes in sourceless SD-brane
387: backgrounds}\label{subsection:buchel}
388:
389: The first progress towards the goal of singularity resolution in
390: SD-brane systems was made by Buchel, Langfelder and Walcher
391: \cite{buchel}.
392: % type %
393: We now briefly review what is, for our purposes, the most relevant
394: point of their work.
395:
396: Essentially, they take the reasonable point of view that the
397: process of creation and subsequent decay of a SD$p$-brane must be
398: driven by a single open string mode: the tachyon field denoted
399: $T(t)$, which lives on the associated unstable D$(p{+}1)$-brane.
400: They use the $p=0$ {\em non}-dilatonic version of the worldvolume
401: % type %
402: action \cite{add1,add2,add3,add4,sen2,sen1}
403: %
404: \beq \label{sourcebuc}
405: S_{brane} = -T_{p+1} \int d^{p+2}y \, e^{-\Phi} V(T) \sqrt{-det \,
406: \left( {\cal P
407: }G_{\alpha\beta}+\partial_{\alpha}T\partial_{\beta}T \right)} +
408: \mu_{p+1}\int f(T)dT\wedge C^{(p+1)}, \eeq
409: %
410: to study the dynamics of the tachyon and its couplings with bulk
411: (closed string) modes. The operation ${\cal P}$ is for pullback, and
412: $\alpha,\beta=1\ldots(p{+}1)$. In section \ref{section:prelim} we will
413: comment both on the validity of this type of effective action, and on
414: the expected form for the tachyon potential $V(T)$ and the
415: Ramond-Ramond coupling $f(T)$. For now, we just use it.
416:
417: The way we look at the calculation of ref.~\cite{buchel} is as
418: follows. Supergravity SD-brane fields should be regarded as
419: arising directly from a large number of unstable branes. Then,
420: using the intuition gained from studying the enhan{\c{c}}on
421: mechanism \cite{enhanconjpp}, it is natural to use an unstable
422: brane probe to study more substantively the candidate supergravity
423: solutions of refs.~\cite{gutperle2,KMP}. Then, we look for
424: problems arising in the probe calculation. The idea is that
425: whenever the probe analysis goes wrong, it signals a pathology for
426: the gravitational background. There are at least two ways the
427: probe analysis can signal a problem: infinite energy or pressure
428: density for the tachyon may be induced, ($\rho_{probe}, p_{probe}
429: \rightarrow \pm \infty$), or there might not exist any reasonable
430: solutions for $T(t)$ across the horizon.
431:
432: So let us consider inserting an unstable brane probe in a
433: background with fields corresponding to the sourceless SD-brane
434: supergravity fields, eqs.~(\ref{bigmetric}), of the previous
435: subsection. The equation of motion for the open string tachyon
436: is, generally,
437: %
438: \begin{eqnarray}\label{eomt_one}
439: && (-g_{tt})^{\frac{1}{2}} (g_{y^{i}y^{i}})^{\frac{p}{2}}
440: (g_{y^{1}y^{1}})^{\frac{1}{2}} e^{-\Phi}
441: \frac{\partial V(T)}{\partial
442: T}\Delta^{\frac{1}{2}} \nonumber\\
443: + && f(T)\, F^{(p+2)} + \frac{d}{dt}\left( \dot{T}\, V(T)
444: \frac{(g_{y^{1}y^{1}})^{\frac{1}{2}}(g_{y^{i}y^{i}})^{\frac{p}{2}}}
445: {e^{\Phi}\Delta^{\frac{1}{2}} (-g_{tt})^{\frac{1}{2}}}\right)=0,
446: \end{eqnarray}
447: %
448: where ${\dot{\ }}\equiv d/dt$, our notation for the Ramond-Ramond
449: field strength is $F^{(p+2)}=dC^{(p+1)}$, we have factored out
450: $T_{p+1}$ by including a factor of $g_s$ in $C^{(p+1)}$, and we
451: also defined the following expression, \beq \Delta = 1 +
452: \frac{(\dot{T})^{2}}{g_{tt}}.\eeq
453:
454: The question needing attention here is whether or not the tachyon
455: field, regarded as a probe,\footnote{The unstable brane will be a
456: probe as long as its backreaction is small and can therefore be
457: treated self-consistently as a perturbation.} is well-behaved when
458: inserted in the candidate supergravity backgounds
459: (\ref{bigmetric}). Ref.~\cite{buchel} provided a clear answer for
460: the case of a $d=4$ SD0-brane with dilaton field set to
461: %
462: zero.\footnote{The authors of ref.~\cite{buchel} also investigated
463: the effect of tachyon backreaction, but the ansatz they used for
464: the supergravity fields was not general enough to handle our cases
465: of interest. In particular, our general equations {\em do not}
466: reduce to theirs upon consistent truncation. Also, their
467: exposition of their backreaction analysis was {\em extremely}
468: brief.}
469: %
470: Let us now see how this goes.
471:
472: The $d=4$ SD0-brane background introduced in ref.~\cite{gutperle}
473: has the form, \beq g_{tt} = -g^{y^{1}y^{1}} =
474: -\frac{Q^{2}}{\omega^{2}} \frac{t^{2}}{t^{2}-\omega^{2}}, \;\;\;\;
475: g_{y^{i}y^{i}}=0, \;\;\;\; g_{xx} = \frac{Q^{2}}{\omega^{2}}t^{2},
476: \eeq and \beq F_{2} = Q \epsilon_{2} \,,\quad \Phi=0. \eeq This
477: spacetime metric has a regular horizon (a coordinate singularity)
478: at $t=\omega$, and a genuine timelike curvature singularity at
479: $t=0$. The expressions for the energy density ($T^{t}_{t}$) and
480: the pressure ($T^{y^{1}}_{y^{1}}$) associated with the probe are
481: respectively, \beq \rho_{probe} \sim
482: \frac{V(T)}{\Delta^{1/2}}\,,\quad p_{probe} \sim V(T)
483: \Delta^{1/2}. \eeq
484:
485: The only time (apart from $t=0$) when the probe limit becomes
486: ill-defined is around $t=\omega$. In fact, in the near horizon
487: limit, the dynamical quantity $\Delta$ satisfies the simple
488: ordinary differential equation \beq \Delta^{2} - \Delta +
489: (t-\omega)\dot{\Delta} = 0. \eeq This has the general solution
490: \beq \Delta = \frac{t-\omega}{t-\omega + g}, \eeq where $g$ is a
491: constant of integration. There are two possible solutions at
492: $t=\omega$: $\Delta=0$ ($g\neq 0$) or $\Delta =1$ ($g=0$).
493: Clearly, the case for which $\Delta =0$ corresponds to the probe
494: limit breaking down since the energy density of the unstable brane
495: diverges. The other possibility, $\Delta=1$, implies that the
496: time-derivative of the tachyon diverges on the horizon. This last
497: case is clearly pathological and cannot correspond to a physically
498: relevant tachyon field solution.
499:
500: The conclusion is that unstable brane {\em probes} are not
501: well-defined in the SD0-brane background. Not only are they useless to
502: resolve the timelike singularity at $t=0$ but, worse, they appear to
503: generically induce a spacelike curvature singularity on the horizon at
504: $t=\omega$. That is, if we take the probe story to be a good
505: indicator of the story for the full backreacted problem.
506:
507: One of our first motivations for the work leading to our paper was
508: to plumb how restricted the conclusions of ref.~\cite{buchel}
509: were. Did this above story work only for bulk couplings of the
510: kind arising for SD0-branes in $d=4$, where no dilaton field
511: appears? Was it true only for the case of SD0-branes, which are a
512: special case for SD$p$-branes since there can be no anistropy in a
513: one-dimensional worldvolume? Are all timelike clothed
514: singularities turned into naked spacelike singularities by probes?
515: Could we even trust the probe approximation to tell us anything
516: about the solution with full backreaction?
517:
518: The first generalization we considered was to look at an unstable
519: brane probe in the background of the isotropic SD$p$-brane
520: solutions of ref.~\cite{KMP}. However, what we saw there was that
521: the naked spacelike singularities remained naked spacelike
522: singularities; the small effect of the probe could not undo that
523: pathology. Next, we moved to analyzing anisotropic solutions of
524: the form (\ref{bigmetric}), those with regular horizons. Some of
525: these are actually completely nonsingular; we analyzed the details
526: of the probe computation in those backgrounds, and the specifics
527: are recorded in Appendix \ref{regular}. The results there are
528: simple to summarize: the solutions with singularities hidden
529: behind horizons do not give rise to conclusions qualitatively
530: different than what we have reviewed here for $d=4$ SD0-branes.
531: The picture therefore remains unsatisfactory.
532:
533: The upshot, then, is that the probe story does not resolve
534: singularities found for sourceless supergravity SD-brane solutions.
535: So we now move to the full backreacted problem for SD$p$-branes in
536: $d{=}10$, which is the main content of our paper.
537:
538: %====================================================================+
539: \section{Supergravity SD-branes with a tachyon
540: source}\label{section:newsetup}
541:
542: In this section we find, in the context of supergravity, the
543: equations of motion associated with the real-time (formation and)
544: decay of a clump of unstable D-branes. We begin this section by
545: writing the form of the action which we will use in our analysis.
546: We will concentrate on only the most relevant\footnote{Relevant in
547: the technical sense.} modes in both the open and closed string
548: sectors; in other words, we keep in our analysis only the
549: (homogeneous) tachyon and massless bulk supergravity fields.
550: Potentially, the effect of massive modes could be encoded in a
551: modified equation of state for the tachyon fluid on the unstable
552: D-brane.\footnote{We thank Andy Strominger for this suggestion.}
553: We leave for future work the issue of non-homogeneous tachyonic
554: modes, and of massive string modes in both the open and closed
555: string sectors, for the coupled bulk-brane system with full
556: backreaction. In order to use the supergravity approximation here
557: self-consistently, we will take $g_s$ small but $g_s N$ large, and
558: time-derivatives will be small compared to $\ell_s$. We will see
559: that it is simple to choose boundary conditions in our numerical
560: integration such that these remain true for all time.
561:
562: %--------------------------------------------------------------------+
563: \subsection{Preliminaries: action and equations of motion}
564: \label{section:prelim}
565:
566: For this section we will be able to suppress R-R Chern-Simons
567: terms in writing the bulk action. This is a consistent
568: truncation, to set the NS-NS two-form to zero throughout the
569: evolution of the system of interest, as long as consistency
570: conditions on the R-R fields are satisfied. {\it E.g.} for the
571: SD2-brane system with R-R field $C^{(3)}$ activated, it is
572: necessary to make certain that $dC^{(3)}\wedge dC^{(3)}=0$ in
573: order not to activate the NS-NS two-form and the accompanying
574: Chern-Simons terms. Other cases are related to this one by
575: T-duality. Therefore, we allow only electric-type coupling of the
576: SDp-brane to $C^{(p+1)}$ (or equivalently magnetic-type coupling
577: to $C^{(7-p)}$). Later we will show that this ansatz is
578: physically consistent provided we assume that there is $ISO(p+1)$
579: symmetry along the worldvolume of the SDp-brane, the object we are
580: interested in. This is equivalent to considering {\em only the
581: lowest-mass tachyon}, {\it i.e.}, not allowing any excitations of
582: the brane tachyon along the spatial worldvolume directions. Of
583: course, the R-R field strengths are then very simple:
584: $G^{(p+2)}=dC^{(p+1)}$, and the string frame bulk action takes the
585: form \cite{polchinski}
586: %
587: \begin{equation}\label{bulk_action}
588: S_{{\rm{bulk}}}={\frac{1}{1 6\pi G_{1 0}}} \int d^{1
589: 0}x\,\sqrt{-g}\left\{ e^{-2\Phi}\left[{\cal
590: R}+4(\partial\Phi)^{2}\right]
591: -{\frac{1}{2}}\left|dC^{p+1}\right|^{2} \right\} \, ,
592: \end{equation}
593: %
594: where ${\cal R}$ is the Ricci scalar. We use a mostly plus
595: signature. In the above conventions, the R-R field solutions
596: automatically get a factor of $g_s$ (as we mentioned also in the
597: previous subsection), and
598: %
599: \begin{equation}\label{Newton}
600: 16\pi G_{10}= (2\pi)^{7}g_{s}^{2}\ell_{s}^{8}\,,\quad
601: \tau_{Dp}={\frac{1}{g_{s}(2\pi)^{p}\ell_{s}^{p+1}}}.
602: \end{equation}
603: %
604:
605: Our analytical and numerical results in following sections will be
606: given in {\em string frame}. If desired, it is easy to convert to
607: Einstein frame -- with metric ${\tilde{g}}_{\mu\nu}$ -- with
608: canonical normalization of the metric and positive dilaton kinetic
609: energy, by using the standard $d=10$ conformal transformation
610: %
611: \begin{equation}\label{string_Einstein}
612: {\tilde{g}}_{\mu\nu} = e^{-(\Phi-\Phi_\infty)/2}g_{\mu\nu}.
613: \end{equation}
614: %
615: Stress-tensors are defined in Einstein frame,
616: %
617: \begin{equation}\label{stress_definition}
618: {\tilde{T}}_{\mu\nu}\equiv{\frac{-1}{\sqrt{-{\tilde{g}}}}}
619: {\frac{\delta S_{{\rm{matter}}}}{\delta {\tilde{g}}^{\mu\nu}}},
620: \end{equation}
621: %
622: with the usual
623: %
624: \begin{eqnarray}\label{stress_tensor_CT}
625: {\tilde{T}}_{\mu\nu}\left[\Phi\right]&=&{\frac{1}{2}}\left[
626: \partial_{\mu}\Phi\partial_{\nu}\Phi
627: -{\frac{1}{2}}{\tilde{g}}_{\mu\nu}
628: ({\tilde{\partial\Phi}})^{2}\right]\, , \\
629: %
630: {\tilde{T}}_{\mu\nu}\left[C^{p+1}\right] &=& {\frac{1}{2(p+1)!}}
631: e^{(3-p)\Phi/2} \left[ {\tilde{G}}_{\mu}^{\
632: \lambda_{2}\ldots\lambda_{p+2}
633: }G_{\nu\lambda_{2}\ldots\lambda_{p+2}} - {\frac{1}{2(p+2)}}
634: {\tilde{g}}_{\mu\nu}{\tilde{G}}^{2} \right].
635: \end{eqnarray}
636: %
637: We can transform to string-frame ``Einstein'' equation using
638: standard formul\ae\
639: %
640: \begin{eqnarray}\label{Einstein_string}
641: && {\tilde{{\cal R}}}_{\mu\nu} -
642: {\frac{1}{2}}{\tilde{g}}_{\mu\nu}{\tilde{{\cal R}}}
643: -{\frac{1}{2}}\left[\partial_{\mu}\Phi\partial_{\nu}\Phi
644: -{\frac{1}{2}}{\tilde{g}}_{\mu\nu}({\tilde{\partial\Phi}})^{2}
645: \right] \\
646: %
647: = &&{\cal R}_{\mu\nu} - {\frac{1}{2}}g_{\mu\nu}{\cal R} +
648: 2\left[\nabla_{\mu}\partial_{\nu}\Phi - g_{\mu\nu}\nabla^{2}\Phi +
649: g_{\mu\nu}(\partial\Phi)^{2}\right]. \nonumber
650: \end{eqnarray}
651: %
652: For all matter fields except the dilaton, it is therefore obvious
653: that string frame ``stress-tensors'' take the form
654: %
655: \begin{equation}\label{peasy_string_stress}
656: T_{\mu\nu} = {\frac{-1}{\sqrt{-g}}} {\frac{\delta S_{\rm
657: matter}}{\delta g^{\mu\nu}}}.
658: \end{equation}
659: %
660: For the dilaton we find
661: %
662: \begin{equation}\label{stress_string_dilaton}
663: T^{\mu}_{\ \nu}\left[\Phi \right] =
664: 2\left[-\nabla^{\mu}\partial_{\nu}\Phi + g^{\mu}_{\ \nu}
665: \nabla^{2}\Phi-g^{\mu}_{\ \nu}(\partial\Phi)^{2} \right] \, ,
666: %
667: \end{equation}
668: %
669: and, obviously, familiar energy conditions for bulk fields are only
670: satisfied in Einstein frame, not string frame.
671:
672: For the bulk field equations, we must include coupling to the brane
673: tachyon --- this is of course an important point of our paper. Hence,
674: we now turn to the brane action. The brane theory is that appropriate
675: to unstable D($p+1$)-branes, for SD$p$-branes. We consider $N$ branes.
676: At low energy, the overall $U(1)$ center-of-mass tachyon $T$ couples
677: as follows:
678: %
679: \begin{equation}\label{brane_worldvolume_action}
680: S_{{\rm{brane}}} = {\frac{\Lambda}{16\pi G_{10}}} \left\{
681: \int dt d^{p+1}y \left[ - \ e^{-\Phi} \sqrt{-A}V(T)\right]
682: + \int f(T)dT\wedge C^{p+1} \right\} \, ,
683: \end{equation}
684: %
685: where the matrix $A_{\alpha\beta}$ is defined as
686: %
687: \begin{equation}\label{A_definition}
688: A_{\alpha\beta} = {\cal{P}}\left(G_{\alpha\beta}+B_{\alpha\beta}
689: \right) +F_{\alpha\beta} +
690: \partial_{\alpha}T\partial_{\beta}T \, ,
691: \end{equation}
692: where ${\cal{P}}$ stands for pullback. For the constants, our
693: conventions are that $T$ is normalized like $F_{\alpha\beta}$,
694: and also
695: \begin{equation}\label{constant_convention}
696: \Lambda\equiv {\frac{N\mu_{p+1}}{g_{s}}} (16\pi G_{10}) =
697: (Ng_{s})(2\pi\ell_s)^{6-p}.
698: \end{equation}
699: Notice that $\Lambda$ is proportional to $g_{s}N$. This will be
700: the {\em sole} continuous\footnote{$\Lambda$ is effectively
701: continuous in the supergravity approximation since $g_{s}N$ is
702: large and all derivatives small in string ($\ell_s$) units.}
703: control parameter associated with the physics of our final
704: solutions for the coupled tachyon-supergravity system.
705:
706: It is important to know when we can expect to trust the action we
707: use. Our approach consists in assuming that the kinetic terms of the
708: open string tachyon field are re-summed to take a Born-Infeld--like
709: form. Strictly speaking, this has only be shown to be a valid claim
710: late in the tachyon evolution. We refer the reader to ref.~\cite{sen1}
711: for more details on the limits in which
712: this approximation holds (see also ref.~\cite{kutasov2}).
713: %
714: The functions $V(T)$ and $f(T)$ are therefore not known exactly at
715: all times. For definiteness in our numerical analysis, we will
716: choose a specific form and assume that the dynamics of the tachyon
717: is governed by eq.~(\ref{brane_worldvolume_action}). For a
718: SD-brane, we make the choice $V(T) = 1/\cosh(T/\sqrt{2})=|f(T)|$
719: which has been shown to be the correct large-$T$ behavior of the
720: couplings. Our results turn out to be quite robust, in that their
721: qualitative features do not depend on the precise form of $V(T)$
722: and $f(T)$.
723:
724: For the remainder of our discussion it will be convenient to use
725: static gauge, which is an appropriate gauge choice for our problem
726: of interest. Therefore, in the following, we will be rather
727: cavalier about dropping pullback signs.
728:
729: When we get to solving the coupled brane-bulk equations, it will
730: be convenient to allow for a density of branes, denoted
731: $\rho_{\perp}$, in the transverse space:
732: %
733: \begin{equation}\label{branedensity}
734: \int_{{\rm{brane}}} dt d^{p+1}y \longrightarrow \int_{{\rm{bulk}}}
735: dt d^{p+1}x d^{8-p}x \rho_{\perp} \, .
736: \end{equation}
737: %
738: Therefore our brane action becomes
739: %
740: \begin{equation}\label{brane_action}
741: S_{{\rm{brane}}} = {\frac{\Lambda}{16\pi G_{10}}} \int d^{10}x
742: \rho_{\perp} \left\{ - V(T)\sqrt{-A}e^{-\Phi} + f(T)
743: \epsilon^{\lambda_{1}\ldots\lambda_{p+2}}
744: C_{[\lambda_{2}\ldots\lambda{p+2}} \partial_{\lambda_{1}]}T
745: \right\},
746: \end{equation}
747: %
748: where $\epsilon$ is the worldvolume permutation tensor with values
749: $(0,\pm 1)$.
750:
751: We are now ready to write down the coupled field equations. The
752: simplest bulk equation to pick off is the dilaton. In string frame we
753: see immediately that
754: %
755: \begin{equation}\label{bulk_dilaton}
756: \nabla^{2}\Phi - (\partial\Phi)^{2} + {\frac{1}{4}} {\cal R} =
757: {\frac{1}{8}} \left({\frac{\Lambda\rho_{\perp}}{\sqrt{-g}}}\right)
758: e^{+\Phi}V(T)\sqrt{-A},
759: \end{equation}
760: %
761: and for the Ramond-Ramond field
762: %
763: \begin{equation}\label{bulk_Ramond-Ramond}
764: \nabla_{\mu}G^{\mu\lambda_{2}\ldots\lambda_{p+2}} =
765: -\left({\frac{\Lambda\rho_{\perp}}{\sqrt{-g}}}\right) f(T)
766: \epsilon^{\mu\lambda_{2}\ldots\lambda_{p+2}}\partial_{\mu}T.
767: \end{equation}
768: For the metric equation of motion (string frame ``Einstein''
769: equations), it is convenient to define
770: %
771: \begin{equation}\label{new_einstein}
772: {\cal R}^{\mu}_{\ \nu} = {\hat{T}}^{\mu}_{\ \nu} \equiv T^{\mu}_{\
773: \nu} - {\frac{1}{8}} T^\lambda_\lambda .
774: \end{equation}
775: %
776: Therefore, we have
777: %
778: \begin{equation}\label{hat_dilaton_stress}
779: {\hat{T}}^{\mu}_{\ \nu} [\Phi]= -2\nabla^{\mu} \partial_{\nu} \Phi
780: -{\frac{1}{4}} g^{\mu}_{\ \nu} \nabla^2\Phi + {\frac{1}{2}}
781: (\partial\Phi)^2 g^{\mu}_{\ \nu} .
782: \end{equation}
783: %
784: For the brane stress-tensor, we need to figure out the dependence
785: of $\sqrt{-A}$ on $g_{\mu\nu} $ (the Wess-Zumino term clearly does
786: not contribute). We find
787: %
788: \begin{equation}\label{brane_stress_one}
789: T^{\mu}_{\ \nu}[T] = - {\frac{1}{2}}
790: e^{\Phi}\left({\frac{\Lambda\rho_{\perp}}{\sqrt{-g}}}\right) V(T)
791: \sqrt{-A}(A^{-1})^{\alpha\mu}g_{\alpha\nu} \, ,
792: \end{equation}
793: %
794: \begin{equation}\label{hat_tachyon_stress}
795: {\hat{T}}^{\mu}_{\ \nu} [T] = - {\frac{1}{2}} e^{\Phi}
796: \left({\frac{\Lambda\rho_{\perp}}{\sqrt{-g}}} \right) V(T)
797: \sqrt{-A} \left[(A^{-1})^{\alpha\mu}g_{\alpha\nu} -
798: {\frac{1}{8}}g^{\mu}_{\nu}(A^{-1})^{\lambda\sigma}g_{\lambda\sigma}
799: \right] \, .
800: \end{equation}
801: %
802: The other object we need for the metric equation of motion is
803: %
804: \begin{equation}\label{hat_Ramond_stress}
805: {\hat{T}}^{\mu}_{\ \nu}[C] = {\frac{1}{2(p+1)!}} e^{2\Phi}
806: G^{\mu\lambda\ldots\sigma} G_{\nu\lambda\ldots\sigma}
807: -{\frac{(p+1)}{16}} e^{2\Phi} {\frac{G^{2}}{(p+2)!}} g^{\mu}_{\
808: \nu} \, .
809: \end{equation}
810: %
811: Lastly, for the tachyon we find the equation of motion
812: %
813: \begin{equation}\label{Tachyon_bulk}
814: {\frac{dV}{dT}}e^{-\Phi}\sqrt{-A} - \partial_{\mu}
815: \left[V(T)e^{-\Phi} \sqrt{-A}(A^{-1})^{\mu\alpha}\partial_{\alpha}T
816: \right]
817: + f(T)\epsilon^{\mu\ldots\lambda}G_{\mu\ldots\lambda} =0 \, .
818: \end{equation}
819: %
820: In the Discussion section we will make some remarks about the
821: robustness of these equations.
822:
823: %--------------------------------------------------------------------+
824: \subsection{The homogeneous brane self-consistent
825: ansatz}\label{subsection:specific_ansatz}
826:
827: As pointed out earlier, we are interested in time-dependent
828: processes by which massless Type IIa or Type IIb supergravity
829: fields are sourced by an open-string tachyon mode on the
830: worldvolume of an unstable brane. A reasonable assumption is that
831: the gravitational background generated by backreaction of the
832: rolling tachyon is of the form
833: %
834: \begin{equation}\label{genmet}
835: ds^{2} = -dt^{2} + a(t)^{2} d\Sigma^{2}_{p+1}(k_\parallel) +
836: R(t)^{2} d\Sigma^{2}_{8-p}(k_\perp) \, ,
837: \end{equation}
838: %
839: where the $n$-dimensional Euclidean metric $d\Sigma^2_{n}(k)$ is
840: %
841: \begin{equation} \label{little}
842: d\Sigma^2_{n}(k) =\left\{
843: \begin{array}{ll}
844: %\vphantom{\sum_{i=1}^{n}}
845: d\Omega^2_{n}& {\rm for}\; k = +1\\
846: dE_n^2& {\rm for}\; k = 0 \\
847: %\vphantom{\sum_{i=1}^{n}}
848: d H^2_{n} &{\rm for}\; k = -1\ ,
849: \end{array} \right.
850: \end{equation}
851: %
852: where $d\Omega^2_{n}$ is the unit metric on $S^{n}$, $dE_n^2$ the flat
853: Euclidean metric, and $dH^2_{n}$ the `unit metric' on $n$--dimensional
854: hyperbolic space $H^{n}$. The corresponding symmetry groups are
855: \begin{equation}\label{symgps}
856: \left\{
857: \begin{array}{ll}
858: %\vphantom{\sum_{i=1}^{n}}
859: SO(n+1) & {\rm for}\; k = +1\\
860: ISO(n) & {\rm for}\; k = 0 \\
861: %\vphantom{\sum_{i=1}^{n}}
862: SO(1,n) &{\rm for}\; k = -1\ .
863: \end{array} \right.
864: \end{equation}
865: For $k{=}\pm 1$ we obviously require that $n\ge 2$.
866:
867: We note that the ansatz (\ref{genmet}), for $k_{\perp}=-1$ and
868: $k_{\parallel}=0$, appears, after using an appropriate change of
869: coordinates, to be equivalent to that considered for supergravity
870: SD-branes in ref.~\cite{KMP}. One can show that this is actually
871: {\em not} the case. In order to bring solutions of the form
872: (\ref{genmet}) with $-\infty < t < +\infty$ to the form introduced
873: in ref.~\cite{KMP} we must find a change of coordinates such that
874: \beq dt^{2} = d\tau^{2} \,
875: F(\tau)^{1/2}(\beta(\tau)\alpha(\tau))^{\frac{2}{7-p}}\left(
876: \frac{\beta(\tau)}{\alpha(\tau)}
877: \right)^{(k_{1}+\tilde{k})(p-1)/(7-p)}, \eeq where $-\infty < \tau
878: < +\infty$. It turns out that for all values of the parameters
879: associated with the isotropic supergravity solutions, such a
880: change of coordinates does not exist. However, the main feature of
881: our analysis is that we are allowing for modifications of SD-brane
882: physics in the region close to the spacelike worldvolume (the
883: region around $t=0$ for the system of coordinates we use). It
884: should therefore have been expected that our new solutions are not
885: included in those presented in ref.~\cite{KMP}. However, we do
886: expect the asymptotics to agree.
887:
888: To be physically relevant, solutions should be asymptotically
889: flat. For example, SD-brane gravity solutions will be such that
890: %
891: \begin{equation}
892: \lim_{t\rightarrow \pm\infty} \dot{a}(t) = 0\, , \quad
893: \lim_{t\rightarrow \pm\infty} \dot{R}(t) = 1\, ,
894: \end{equation}
895: %
896: for $k_{\parallel}=0$ and $k_{\perp}=-1$. We will also see in our
897: numerical analysis that only some values of $k_{\perp}$ and
898: $k_{\parallel}$ are allowed. Also, we expect that for the dilaton
899: and R-R field
900: %
901: \begin{equation}
902: \lim_{t\rightarrow\pm\infty} {\dot{C}}(t) =0\,, \quad
903: \lim_{t\rightarrow\pm\infty} {\dot{\Phi}}(t) =0.
904: \end{equation}
905: %
906: By inspection of the tachyon equation of motion (\ref{Tachyon_bulk}),
907: we see that the electric- (or magnetic-) only ansatz referred to at
908: the beginning of this section will be obviously consistent if we only
909: allow worldvolume time-derivatives. This is tantamount to imposing an
910: $ISO(p+1)$ symmetry on the worldvolume. Ref.~\cite{larsen} argues that
911: spatial inhomogeneities of the tachyon field will play an important
912: role in the decay (a view which is also supported, although using a
913: different line of reasoning, by the results of
914: ref.~\cite{hashimoto,cosmotachyonK}). It will be interesting to
915: investigate the full importance of such effects in the context of our
916: effective supergravity analysis. We will include a discussion of the
917: nontrivial issues raised in the Discussion section.
918:
919: It turns out that the equations for the combined bulk-brane
920: evolution in the time-dependent system are complicated enough to
921: require numerical solution. For this reason, we will not be able
922: to accommodate the most natural ansatz\footnote{Strictly speaking,
923: instead of being a delta-function distribution, the more general
924: ansatz for the source should be extended ({\it e.g.} a Gaussian)
925: with its size of the order of the string length.}
926: $\rho_{\perp}=\delta({\vec{x}})$. Instead, we will use the
927: ``smeared'' ansatz also used by Buchel et al. in
928: ref.~\cite{buchel},
929: %
930: \begin{equation}\label{Buchel_smear}
931: \rho_{\perp} =\rho_{0}\sqrt{g_{\perp}} \, .
932: \end{equation}
933: %
934: Note that in this ansatz, the implicit time dependence in the
935: transverse metric components is not varied in producing the
936: equation of motion, rather it is only taken into consideration at
937: the end of the calculation. Also, a smeared brane source does not
938: contribute stress-energy perpendicular to the worldvolume, which
939: is in the directions $t\,,{\vec{y}}$. It should be noted that the
940: effect of using this ansatz will be minimized by using a small
941: value for the density parameter $\rho_{0}$. Of course, the aim
942: when using such an ansatz is to get rid of any brane action
943: dependence on the transverse coordinates ${\vec{x}}$.
944:
945: We should remark that supergravity solutions corresponding to
946: unstable D-brane systems have been found before \cite{ozetal}.
947: Their solutions are time-independent, a feature which might seem
948: rather unreasonable since they are, after all, supposed to
949: describe unstable objects. Typically, these solutions are nakedly
950: singular; there is no horizon. For reasons discussed previously,
951: these solutions would therefore justifiably be regarded with some
952: level of suspicion. Possibly, we should really regard these
953: solutions as fixed-time snap-shots of the unstable brane system
954: during its evolution. They do however reflect one desirable
955: feature: taking into account warping of space in the directions
956: transverse to the unstable branes.
957:
958: What we really want is a sort of hybrid of that approach -- where
959: transverse dependence is the only dependence -- and what we are doing
960: here -- where time dependence is all there is. This is something we
961: postpone to a future investigation; remarks on this will be given in
962: the Discussion section.
963:
964: Let us now get back to the simplified ansatz, and just go ahead
965: and solve it. We are therefore interested in the precise system
966: of {\em ordinary} differential equations for our coupled
967: tachyon-supergravity system. Using the form $C_{12\ldots
968: p+1}\equiv C(t)$ (which is consistent with our ansatz) the
969: equation of motion for the R-R field (\ref{bulk_Ramond-Ramond})
970: becomes
971: %
972: \begin{equation}\label{RR_ansatz}
973: {\ddot{C}}+ {\dot{C}}\left[(8-p){\frac{{\dot{R}}}{R}}
974: -(p+1){\frac{{\dot{a}}}{a}}\right] = \lambda a^{p+1}f(T)
975: {\dot{T}}.
976: \end{equation}
977: %
978:
979: Now let us find the dilaton equation of motion. A useful identity
980: is
981: %
982: \begin{equation}\label{R_identity}
983: {\cal R} = 5(\partial\Phi)^2 - {\frac{9}{2}}\nabla^2\Phi +
984: {\frac{(3-p)}{8(p+2)!}} e^{2\Phi}G^2 + {\frac{1}{8}}\left(
985: {\frac{\Lambda\rho_\perp}{\sqrt{-g}}} \right)
986: e^{\Phi}V(T)\sqrt{-A}(A^{-1})^{\lambda\sigma}g_{\lambda\sigma} \,
987: ,
988: \end{equation}
989: %
990: with which the dilaton equation of motion can be written,
991: %
992: \begin{equation}
993: 2(\partial\Phi)^2 - \nabla^2\Phi = {\frac{(p-3)}{4(p+2)!}}
994: e^{2\Phi}G^2 + \left( {\frac{\Lambda\rho_\perp}{\sqrt{-g}}}
995: \right) e^{\Phi}V(T)\sqrt{-A} \left[ 1-{\frac{1}{4}}
996: (A^{-1})^{\lambda\sigma}g_{\lambda\sigma} \right] \, .
997: \end{equation}
998: %
999: This last expression is simply the Einstein frame equation of
1000: motion. So the dilaton in our ansatz satisfies
1001: %
1002: \begin{eqnarray}\label{final_dilaton}
1003: && {\ddot{\Phi}} +{\dot{\Phi}}
1004: \left[(8-p){\frac{{\dot{R}}}{R}}+(p+1){\frac{{\dot{a}}}{a}}
1005: \right] -2{\dot{\Phi}}^{2} \nonumber\\ && = {\frac{(3-p)}{4}}
1006: \left({\frac{e^\Phi {\dot{C}}}{a^{(p+1)}}}\right)^2
1007: +{\frac{\lambda}{4}} e^{\Phi} V(T) \left[(3-p)\sqrt{\Delta} -
1008: {\frac{1}{\sqrt{\Delta}}} \right] \, .
1009: \end{eqnarray}
1010: %
1011: This will be used %in the Einstein equations
1012: %
1013: whenever double time-derivatives of the dilaton need to be
1014: substituted for.
1015:
1016: With $T=T(t)$ we find that the dynamics of the tachyon field is
1017: governed by
1018: %
1019: \begin{equation}\label{Tachyon_motion}
1020: {\ddot{T}} = \Delta \left\{ {\dot{\Phi}}{\dot{T}} -
1021: {\dot{T}}\left[ (p+1){\frac{\dot{a}}{a}} \right] -
1022: {\frac{1}{V(T)}}{\frac{dV(T)}{dT}} + {\frac{f(T)}{V(T)}} {\dot{C}}
1023: e^{\Phi} a^{-(p+1)}\sqrt{\Delta} \right\} \, ,
1024: \end{equation}
1025: %
1026: where $\Delta = 1 - \dot{T}^{2}$. In this paper we will be
1027: assuming that $|f(T)|=V(T)$, a statement which has been shown to
1028: be correct only at past and future asymptopia. However, we have
1029: also done numerical experiments which show that some breaking of
1030: this relation at intermediate times (near the hilltop) does not
1031: change the important features of our solutions.
1032:
1033: We now turn to the equations of motion for the metric components
1034: $a(t)$ and $R(t)$. For the stress-tensors, eliminating second
1035: order derivatives in matter fields, we have
1036: %
1037: \begin{eqnarray}\label{stressed_out}
1038: % time dirn
1039: {\hat{T}}^{t}_{\ t} &\! = &\!
1040: 4({\dot{\Phi}})^{2} -2{\dot{\Phi}} \left[
1041: (p+1){\frac{{\dot{a}}}{a}} + (8-p){\frac{{\dot{R}}}{R}} \right]
1042: + {\frac{(5-2p)}{4}}
1043: \left({\frac{e^{\Phi}{\dot{C}}}{a^{p+1}}}\right)^{2}
1044: \cr && \qquad\qquad
1045: + {\frac{1}{4}}\lambda e^{\Phi} V(T)
1046: \left[(7-2p)\sqrt{\Delta}-{\frac{4}{\sqrt{\Delta}}} \right] \, ,
1047: \cr
1048: % parallel dirn
1049: {\hat{T}}^{y}_{\ y} &\! = &\!
1050: 2{\dot{\Phi}} {\frac{{\dot{a}}}{a}}
1051: - {\frac{1}{4}}
1052: \left({\frac{e^{\Phi}{\dot{C}}}{a^{p+1}}}\right)^{2}
1053: - {\frac{1}{4}}\lambda e^{\Phi} V(T) \sqrt{\Delta} \, , \cr
1054: % transverse dirn
1055: {\hat{T}}^{x}_{\ x} &\! = &\! 2{\dot{\Phi}} {\frac{{\dot{R}}}{R}}
1056: + {\frac{1}{4}}
1057: \left({\frac{e^{\Phi}{\dot{C}}}{a^{p+1}}}\right)^{2}
1058: + {\frac{1}{4}}\lambda e^{\Phi} V(T) \sqrt{\Delta} \, .
1059: \end{eqnarray}
1060: %
1061: The components of the Ricci tensor are easily
1062: evaluated:
1063: %
1064: \begin{eqnarray}\label{Ricci_components}
1065: {\cal R}^{t}_{\ t}&=& (p+1){\frac{\ddot{a}}{a}}
1066: +(8-p){\frac{\ddot{R}}{R}} \, , \\
1067: %
1068: {\cal R}^{y}_{\ y} &=& {\frac{\ddot{a}}{a}} +
1069: (8-p){\frac{\dot{a}}{a}}{\frac{\dot{R}}{R}} + p\left[
1070: \left({\frac{\dot{a}}{a}}\right)^2
1071: + {\frac{k_\parallel}{a^2}} \right] \, , \\
1072: %
1073: {\cal R}^{x}_{\ x}&=&{\frac{\ddot{R}}{R}} +
1074: (p+1){\frac{\dot{a}}{a}}{\frac{\dot{R}}{R}} +
1075: (7-p)\left[\left({\frac{\dot{R}}{R}}\right)^{2}
1076: +{\frac{k_{\perp}}{R^{2}}}\right] \, .
1077: \end{eqnarray}
1078: %
1079: For the (string-frame) ``Einstein'' equations, the time,
1080: longitudinal and transverse components are respectively
1081: %
1082: \begin{eqnarray} \label{magic} (p+1){\frac{\ddot{a}}{a}}
1083: +(8-p){\frac{\ddot{R}}{R}} &= & + 4({\dot{\Phi}})^{2}
1084: -2{\dot{\Phi}} \left[
1085: (p+1){\frac{{\dot{a}}}{a}} + (8-p){\frac{{\dot{R}}}{R}} \right]
1086: \cr &&
1087: + {\frac{(5-2p)}{4}}
1088: \left({\frac{e^{\Phi}{\dot{C}}}{a^{p+1}}}\right)^{2}
1089: + {\frac{1}{4}}\lambda e^{\Phi} V(T)
1090: \left[(7-2p)\sqrt{\Delta}-{\frac{4}{\sqrt{\Delta}}} \right] \,
1091: , \end{eqnarray}
1092: %
1093: \beq \label{aeq} {\frac{\ddot{a}}{a}} = -
1094: (8-p){\frac{\dot{a}}{a}}{\frac{\dot{R}}{R}} -
1095: p\left[\left({\frac{\dot{a}}{a}}\right)^2
1096: + {\frac{k_\parallel}{a^2}} \right]
1097: + 2{\dot{\Phi}} {\frac{{\dot{a}}}{a}} - {\frac{1}{4}}
1098: \left({\frac{e^{\Phi}{\dot{C}}}{a^{p+1}}}\right)^{2}
1099: - {\frac{1}{4}}\lambda e^{\Phi} V(T) \sqrt{\Delta} \, , \eeq
1100: %
1101: \beq \label{req} {\frac{\ddot{R}}{R}} = -
1102: (p+1){\frac{\dot{R}}{R}}{\frac{\dot{a}}{a}} -
1103: (7-p)\left[\left({\frac{\dot{R}}{R}}\right)^{2}
1104: +{\frac{k_{\perp}}{R^{2}}}\right]
1105: + 2{\dot{\Phi}} {\frac{{\dot{R}}}{R}} + {\frac{1}{4}}
1106: \left({\frac{e^{\Phi}{\dot{C}}}{a^{p+1}}}\right)^{2}
1107: + {\frac{1}{4}}\lambda e^{\Phi} V(T) \sqrt{\Delta} \, . \eeq
1108: %
1109: %
1110: In the end we have a system of five second order coupled ordinary
1111: differential equations for $\{T,C,\Phi,a,R\}$ as functions of $t$.
1112: These are respectively equations (\ref{Tachyon_motion}),
1113: (\ref{RR_ansatz}), (\ref{final_dilaton}), (\ref{aeq}) and
1114: (\ref{req}). For consistency this system of equations must be
1115: supplemented with the first-order constraint
1116: %
1117: \beq\begin{array}{l} \label{fconstraint}
1118: {\displaystyle{
1119: \frac{1}{\lambda e^{\Phi}}\left\{
1120: 2(p+1)(8-p)\frac{\dot{a}}{a}\frac{\dot{R}}{R} + p(p+1) \left[
1121: \left(\frac{\dot{a}}{a}\right)^{2}+\frac{k_{\parallel}}{a^{2}}\right]
1122: +(7-p)(8-p)\left[
1123: \left(\frac{\dot{R}}{R}\right)^{2}+\frac{k_{\perp}}{R^{2}}\right]
1124: \right. }}
1125: \nonumber \\ \qquad\qquad {\displaystyle{ \left. -4\dot{\phi}\left[
1126: (p+1)\frac{\dot{a}}{a}+(8-p)\frac{\dot{R}}{R}-\dot{\phi}\right]-\frac{1}{2}
1127: \left(\frac{e^{\Phi}\dot{C}}{a^{p+1}}\right)^{2}\right\} }}
1128: ={\displaystyle{
1129: \frac{V(T)}{\sqrt{\Delta}} }}
1130: \end{array}\eeq
1131: %
1132: obtained by plugging eqs.~(\ref{aeq}) and (\ref{req}) in
1133: eq.~(\ref{magic}). Of course, if this last equation is satisfied at
1134: $t=0$ it will be for all times.
1135:
1136: %====================================================================+
1137: \section{The roll of the tachyon}
1138: \label{section:numeric}
1139:
1140: We refer to a supergravity SD-brane as the field configuration
1141: generated by a system composed of a large number, $N$, of
1142: microscopic SD-branes. As emphasized in section
1143: \ref{subsection:specific_ansatz}, there is a single continuous
1144: parameter that controls the dynamics of these fields, {\it i.e.},
1145: $\lambda = \rho_{0} \Lambda$. This is the parameter that
1146: determines the relative importance of the unstable brane source.
1147: Clearly, for $\lambda\rightarrow 0$ the open string tachyon
1148: decouples from the bulk fields (no backreaction) and the
1149: corresponding supergravity solutions will presumably be the
1150: singular ones found in refs.~\cite{gutperle,gutperle2,KMP}.
1151:
1152: In this section we present solutions with the symmetries of
1153: SD-branes and a non-vanishing $\lambda$. We demonstrate that they
1154: are generically non-singular.
1155:
1156: %--------------------------------------------------------------------+
1157: \subsection{Tachyon evolution in flat space}
1158:
1159: The solutions relevant for SD-brane physics in Type II a,b
1160: superstring theory could be the ones corresponding to an open
1161: string tachyon evolving from \beq \label{bc5} \lim_{t\rightarrow
1162: -\infty} T(t) = +\infty \eeq to \beq \lim_{t\rightarrow +\infty}
1163: T(t) = -\infty \eeq in a symmetric runaway potential of the form
1164: shown on figure \ref{potential}. Solutions of this type correspond
1165: to the tachyon evolving between two different minima of the
1166: potential. Possible initial conditions (at $t=0$) for these
1167: solutions are of the form \beq \label{bc7} \dot{T}(0) = {\rm
1168: const.} \, , \;\;\;\;\;\; T(0)=0 \, . \eeq Another set of
1169: solutions corresponds to a tachyon evolution with \beq \label{frg}
1170: \lim_{t\rightarrow \pm\infty} T(t) = \alpha \, \infty \, , \eeq
1171: where $\alpha=1$ and $\alpha=-1$ lead to equivalent solutions.
1172: Appropriate boundary conditions for the tachyon (at $t=0$) are
1173: then of the form \beq \label{bc8} \dot{T}(0) = 0 \, , \;\;\;\;\;\;
1174: T(0)= {\rm const.} \eeq In this section we characterize the
1175: supergravity solutions generated by a tachyonic source with the
1176: properties mentioned above. The results we present are for
1177: solutions with boundary conditions (\ref{bc8}) and asymptotic
1178: behavior (\ref{frg}). We have found that the solutions associated
1179: with the boundary conditions (\ref{bc7}) and asymptotic behavior
1180: (\ref{bc5}) have similar qualitative features. Strictly speaking,
1181: our approach to studying real-time evolution in supergravity can
1182: also be extended to more general cases, {\it i.e.}, brane decay or
1183: creation (half SD-branes). \FIGURE{\epsfig{file=potential.eps,
1184: width = 10cm}\caption{Possible form of the open string tachyon
1185: potential $V(T)$ on a SD-brane in Type II a,b superstring theory.}
1186: \label{potential}}
1187:
1188: In our analysis we use the potential \beq \label{pot12}
1189: V(T)=\frac{1}{\cosh \left(T/\sqrt{2}\right)}\eeq because it agrees
1190: with open string field theory calculations for large values of the
1191: tachyon for unstable D-brane systems in Type II a,b superstring
1192: theory.\footnote{In bosonic string theory the potential is
1193: asymmetric and unbounded from below as $T\rightarrow -\infty$.} It
1194: is not known what the exact potential is for intermediate times
1195: but our solutions are only mildly affected by its particular form.
1196: The expression for the tachyon when $\lambda=0$, {\it i.e.}, when
1197: there are no couplings to the bulk supergravity modes,
1198: is\footnote{We refer the reader to Appendix \ref{flatach} for a
1199: derivation of this expression.}
1200: %
1201: \beq \label{flatfield} T(t) = -\sqrt{2}\;{\rm arcsinh} \left(
1202: -\dot{T}(0) \sinh \frac{t}{\sqrt{2}} \right) \eeq
1203: %
1204: for the boundary conditions (\ref{bc7}). When the boundary
1205: conditions are of the form (\ref{bc8}), the analytic expression
1206: for the tachyon is \beq \label{flatfield2} T(t) = -\sqrt{2}\;{\rm
1207: arcsinh} \left( \sinh \left( -\frac{T(0)}{\sqrt{2}} \right) \cosh
1208: \frac{t}{\sqrt{2}} \right). \eeq Figure \ref{T} shows a tachyon
1209: profile for $T(0)=0.83$ and $\dot{T}(0)=0$. Homogeneous solutions
1210: such as eqs.~(\ref{flatfield}) and (\ref{flatfield2}), derived
1211: from a tachyonic action, correspond to a fluid that has a constant
1212: energy density and vanishing pressure asymptotically (tachyon
1213: matter). We will shortly see how these features are affected by
1214: the inclusion of couplings to bulk modes.
1215: \FIGURE{\epsfig{file=tach1.eps, width = 10cm}\caption{The tachyon
1216: field evolution for $\dot{T}(0)=0$ and $T(0)=0.83$.} \label{T}}
1217:
1218: \subsection{R-symmetry group}
1219: Before presenting the numerical results we comment on the issue of
1220: R-symmetry. As pointed out in ref.~\cite{gutperle}, SD-brane
1221: solutions should be invariant under the transverse Lorentz
1222: transformations leaving the location of the brane fixed. This
1223: corresponds to an $SO(1,8-p)$ R-symmetry.\footnote{The
1224: interpretation in terms of R-symmetry is relevant for the idea
1225: that SD-brane gravity fields are holographically dual to a
1226: worldvolume Euclidean field theory.} This property is embodied in
1227: the supergravity solutions found in
1228: refs.~\cite{gutperle,gutperle2,KMP} where the transverse space
1229: metric has a factor of the form: $R(t)^{2} dH_{8-p}^{2}$. The
1230: embedding group of the hyperbolic space $H^{8-p}$ is
1231: $SO(1,8-p)$.\footnote{The intuition for the nature of the
1232: R-symmetry group is inherited from the AdS/CFT correspondence. For
1233: example, the metric of a 3-brane is of the form: $ds^{2}=
1234: f(r)\left[ -dt^{2} + d{\vec y}^{2} \right] + 1/f(r)\left[ dr^{2} +
1235: r^{2} d\Omega_{5}^{2} \right]$. The R-symmetry group in this case
1236: is $SO(6)$, a statement that can be traced to the fact that the
1237: near horizon limit of the geometry is AdS$_{5}\times S^{5}$, {\it
1238: i.e.}, the gravitational background dual to the worldvolume field
1239: theory on an ensemble of D3-branes.} This explains why we study
1240: supergravity solutions with $k_{\parallel}=0$ and $k_{\perp}=-1$
1241: in more details.
1242: %
1243: However, the cases with $k_{\perp}=0$ are also candidate solutions
1244: for SD-brane (and, more generally, unstable D-brane) supergravity
1245: solutions. We present an analysis of those and other cases in
1246: section \ref{othercases}.
1247:
1248: It is suggested in ref.~\cite{gutperle} that the spacelike naked
1249: singularities \cite{gutperle,gutperle2,KMP} associated with the
1250: supergravity SD-branes could be resolved by using a metric ansatz
1251: that allows for a breaking of the R-symmetry in the region around
1252: the core of the object ($t=0$). The intuition from the AdS/CFT
1253: correspondence comes from ref.~\cite{nunez} which describes cases
1254: of spontaneously broken R-symmetry. Our ansatz as it is cannot
1255: accommodate such an R-symmetry breaking. Presumably, this would
1256: correspond to a time-dependent process by which the R-symmetry is
1257: broken down to $SO(8-p)$ in a finite region. The closest our
1258: ansatz can come to realizing this scenario is if we consider the
1259: cases with $k_{\perp}=0$. Then, the transverse symmetry group is
1260: $ISO(8-p)$, the compactification of which is $SO(8-p)$. We find
1261: that solutions with $k_{\perp}=0$ are {\it regular} and
1262: asymptotically flat. Moreover, as we will see in
1263: section~\ref{othercases}, a generic feature of the $k_{\perp}=0$
1264: solutions is that the effective string coupling
1265: $g_{s}\exp(\Phi(t))$ vanishes asymptotically. We will also remark
1266: on the cases with $k_\perp$ and/or $k_\parallel$ equal to +1,
1267: which all develop big-crunch singularities in finite time.
1268: %
1269:
1270: %--------------------------------------------------------------------+
1271: \subsection{Numerical SD-brane solutions}
1272:
1273: Our ansatz for describing supergravity SD$p$-brane solutions is
1274: %
1275: \begin{eqnarray} ds^{2}_{Sp} &= & -dt^{2} + a(t)^{2}d{\vec y}^{2}
1276: + R(t)^{2} dH_{8-p}^{2}, \nonumber \\
1277: C & =& C(t)\,,\quad \Phi = \Phi(t),
1278: \end{eqnarray}
1279: %
1280: where we have taken $k_{\parallel}=0$ and $k_\perp=-1$ in accord
1281: with the original paper ref.~\cite{gutperle}. Again, these
1282: supergravity bulk modes will be excited by couplings to an
1283: homogeneous open string tachyon as described in section
1284: \ref{section:prelim}. We consider only homogeneous solutions ({\it
1285: i.e.}, depending only on time). As discussed earlier, it is
1286: possible that inhomogeneities might play an important role in the
1287: creation/decay process \cite{larsen}, but we are postponing this
1288: issue for now.
1289:
1290: The resulting system of coupled differential equations for which
1291: we will find solutions is of course highly non-linear. Among other
1292: things, this implies that it is not possible to extract scaling
1293: behavior for the fields from their equations of motion. We could
1294: therefore expect that the behavior of these solutions depends in a
1295: physically important way on the initial conditions for the field
1296: components involved: $a(t)$, $R(t)$, $C(t)$, $\Phi(t)$ and the
1297: source $T(t)$.
1298: %
1299: We will see that the qualitative behavior of the supergravity
1300: SD-brane solutions actually does not depend very much on these
1301: initial conditions.
1302:
1303: The multiplicity of potential solutions could be large: one
1304: solution would be expected for each set of initial conditions.
1305: Now, for SD-branes the number of solutions is greatly restricted
1306: by the constraint equation (\ref{fconstraint}). It would appear
1307: natural to consider, as candidate solutions for SD-branes, those
1308: which are time-reversal symmetric around $t=0$.\footnote{We thank
1309: Alex Buchel for pointing out that the time-reversal symmetric
1310: solutions are inconsistent. In fact, the constraint equation
1311: (\ref{fconstraint}) cannot be satisfied (for the case
1312: $k_{\parallel}=0$ and $k_{\perp}=-1$) with the boundary
1313: conditions: $\dot{a}(0)=0$, $\dot{R}(0)=0$, $\dot{\phi}(0)=0$,
1314: $\dot{C}(0)=0$ and $T(0)=0$.} In fact, precisely time-symmetric
1315: solutions actually do not exist! That is, unless they have
1316: $k_\parallel=0,k_\perp=+1$, in which case they have R-symmetry
1317: which is markedly in conflict with the proposals of earlier
1318: papers, e.g. \cite{gutperle}. As we said above, these solutions
1319: with wrong R-symmetry also develop a big-crunch singularity in
1320: finite time.
1321: %
1322:
1323: In a sense, motivated by inflationary cosmology, the absence of
1324: precisely time-symmetric solutions is not particularly bothersome
1325: to us. What we mean by this is that it is an extremely fine-tuned
1326: situation to have exactly zero kinetic energy in each of the bulk
1327: fields at $t=0$. Any small quantum fluctuation of a bulk field
1328: takes us away from time-symmetry. Therefore, all solutions which
1329: we will exhibit here will have some kinetic energy in one or more
1330: of the bulk fields at $t=0$. And the kick in bulk field(s) at
1331: $t=0$, required to solve the initial constraint, can actually be
1332: made very small by, {\it e.g.}, choosing $R(0)$ large. Such
1333: initial conditions are quite generic: even a little bit of kick in
1334: $a(t)$ alone will suffice to give completely nonsingular solutions
1335: for the entire time-evolution. The precisely time-symmetric
1336: solutions are however inaccessible; see subsection \ref{trssblar}
1337: for a more detailed exposition of this case.
1338: %
1339:
1340: Another reason why we find a slight bulk asymmetry at $t=0$ reasonable
1341: relates to particle/string production. We have of course neglected
1342: such production in our analysis. It is nonetheless clear that, for a
1343: full SD-brane solution, backreaction will combine with particle
1344: production to make bulk fields naturally {\em a}symmetric. It is only
1345: in the approximation of zero backreaction that time-symmetry is
1346: possible when particle production occurs, but of course that
1347: approximation cannot be self-consistent.
1348: %
1349:
1350: Let us now move to the solution of our problem of interest.
1351: Unfortunately we have been unsuccessful at finding analytical
1352: expressions for the bulk modes and tachyon field when their mutual
1353: coupling is non-vanishing ($\lambda\neq 0$). We therefore resorted to
1354: solving the corresponding system of differential equations
1355: numerically.\footnote{Recall that this is the primary reason why it is
1356: so difficult to include ${\vec{x}}$-dependence in our ansatz.}
1357: %
1358: In what follows we show the results associated with a SD4-brane,
1359: and present some interesting analysis of the effect of varying the
1360: initial conditions. We provide general comments for other
1361: SD-branes with $p<7$ and explain how the SD$p$-branes with $p=7,8$
1362: are different. We comment on the robustness of the solutions to
1363: changes in initial conditions. Throughout our analysis we pay
1364: special attention to the impact of varying the initial condition
1365: $\Phi(0)$ on the dilaton, because this controls the string
1366: coupling close to the hilltop.
1367:
1368: For finding the numerical solutions we fix the initial conditions
1369: at $t=0$, and evolve this data forward in time. Then, we evolve
1370: the same data backward in time from $t=0$. The result is the
1371: solution associated with a full SD-brane, {\it i.e.}, the bulk
1372: fields sourced by the open string tachyon as it evolves from
1373: $t=-\infty$ to $t=+\infty$. In this section we present the results
1374: associated with a SD4-brane with boundary conditions
1375: %
1376: \begin{eqnarray} \label{fullbc} T(0)=0.83 \, , \;\;\;
1377: \dot{T}(0)=0 \, , \;\;\; a(0)=1
1378: \, , \;\;\; \dot{a}(0)= 0.091 \, , \;\;\; R(0)=10 \, , \nonumber
1379: \\ \dot{R}(0) = 0.01 \, , \;\;\; \phi(0)=-1 \, , \;\;\;
1380: \dot{\phi}(0) =0.01 \, , \;\;\; C(0) = 0 \, , \;\;\; \dot{C}(0) =
1381: 0.01 \, , \end{eqnarray}
1382: %
1383: and $\lambda = 0.1\, $. Again, these solutions correspond to the
1384: tachyon rolling up the potential from $T=+\infty$ ($t=-\infty$)
1385: and coming to a halt for $T(0)=0.83$ which corresponds to a
1386: turning point. Then, the tachyon evolves toward the bottom of the
1387: potential at $T=+\infty$ for $t=+\infty$. This type of tachyon
1388: evolution was considered, for example, in
1389: refs.~\cite{maldacena,andylastweek} where they are called full
1390: S-branes. We could of course have chosen different initial
1391: conditions - including some with less time-asymmetry; these are
1392: used just for the sake of illustration.
1393: %
1394:
1395: %- - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - +
1396: \subsubsection{Deformation of the tachyon field}
1397:
1398: Figure~\ref{T} shows the evolution of the tachyon for the
1399: S4-brane. For $\lambda=0$ (no coupling between the closed and the
1400: open string modes), we find \beq \lim_{t\rightarrow \pm \infty}
1401: T(t) = \pm t. \eeq This observation is suggestive that the tachyon
1402: field might play the role of time itself in cosmological models
1403: driven by brane decay (as proposed in ref.~\cite{sen1}). We find
1404: that this asymptotic behavior for the tachyon survives when
1405: couplings to bulk modes are introduced. In fact, for $\lambda\neq
1406: 0$ we find \beq \lim_{t\rightarrow \pm\infty} T(t) = \pm (t +
1407: \kappa_{\scriptscriptstyle \pm T}). \eeq The constants
1408: $\kappa_{\scriptscriptstyle \pm T}$ depend non-trivially on the
1409: boundary conditions at $t=0$ and $p$. For example, as $p$ is
1410: increased $\kappa_{\scriptscriptstyle T}$ decreases. Also, for
1411: large values of $\Phi(0)$ the tachyon deformation from the flat
1412: space case becomes larger. As mentioned in Appendix~\ref{flatach},
1413: for $\lambda=0$ the state of the tachyon for large time
1414: ($t\rightarrow \pm \infty$) is that of a perfect fluid with
1415: constant energy density and vanishing pressure. This is the
1416: so-called tachyon matter. We find that for $\lambda\neq 0$, both
1417: the energy density and the pressure (physical quantities measured
1418: in the Einstein frame) vanish. In other words, the tachyon matter
1419: is clearly only an illusion of the $g_{s}\rightarrow 0$ limit.
1420:
1421: We consider briefly the effect of varying the initial conditions
1422: $\dot{T}(0)$ and $T(0)$. For half SD-branes ({\it i.e.}, the
1423: future of the full SD-branes) we find that the time it takes the
1424: tachyon to reach the bottom of its potential increases for smaller
1425: values of $\dot{T}(0)$ and $T(0)$. Not only that, for very small
1426: initial velocities the tachyon stays perched at the top of the
1427: potential for a certain period of time. In general, we observe
1428: that it takes less time for the tachyon to reach the bottom of its
1429: potential when we increase $\lambda$. Now, we also observe that
1430: the difference between the curves associated with flat space
1431: ($\lambda =0$) and $\lambda\neq 0$ tachyons decreases as
1432: $\dot{T}(0)$ and $T(0)$ increase. Also, for large negative values
1433: of $\Phi(0)$, $\kappa_{\scriptscriptstyle T}$ becomes very small.
1434: This is simply a reflection of the fact that such cases correspond
1435: to a very small initial string coupling (see below for details).
1436:
1437: There is another interesting feature of the tachyon when coupled
1438: to the massless closed string modes. Firstly, the time it takes
1439: the tachyon to reach the bottom of its potential is not
1440: significantly altered even when considering large values of the
1441: `coupling' $\lambda$. Finally, we find that as the coupling
1442: $\lambda$ is increased, $\kappa_{\scriptscriptstyle T}$, or the
1443: deformation away from the flat space tachyon, increases
1444: correspondingly.
1445:
1446: %- - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - +
1447: \subsubsection{Time-dependent string coupling}\label{subsec:dildil}
1448:
1449: The string coupling is given by the expression \beq g_{s} =
1450: e^{<\Phi_{0}>} \,, \eeq where $\Phi_{0}$ is the background dilaton
1451: field in the absence of sources, {\it i.e.}, strings and D-branes.
1452: Typically the presence of stringy excitations will modify the
1453: coupling of the theory, \beq g_{s} \rightarrow g_{s}
1454: e^{\Phi(\eta)}, \eeq where $\eta$ is a spacelike variable for
1455: D-branes and is timelike for SD-branes.
1456:
1457: For supergravity D$p$-brane solutions the dilaton field is (see,
1458: for example, ref.~\cite{peet}), \beq g_{s}(r) = g_{s} e^{\Phi(r)}
1459: = g_{s} \left( 1+ \frac{c_{p}g_{s}N_{p}l_{s}^{7-p}}{r^{7-p}}
1460: \right)^{\frac{1}{4}(3-p)}, \eeq where
1461: $c_{p}=(2\sqrt{\pi})^{5-p}\Gamma[\frac{1}{2}(7-p)]$. The string
1462: coupling is seen to vary according to whether test closed strings
1463: propagate close or far from the horizon ($r=0$) of these
1464: geometries. For all static supergravity solutions (including
1465: NS5-branes), the asymptotic string coupling is such that\beq
1466: \lim_{r\rightarrow +\infty} g_{s}(r) = g_{s}. \eeq The effect of
1467: supergravity D-branes is therefore to modify the coupling locally.
1468: For $p<3$, it is large close to the horizon and decreases to
1469: $g_{s}$ as $r\rightarrow +\infty$. For $p>3$, the string coupling
1470: is small close to the horizon but increases to $g_{s}$ for large
1471: $r$. The case $p=3$ is special because the dilaton field sourced
1472: by the 3-brane is constant throughout the spacetime. Typically,
1473: the size of the region where the dilaton is not constant depends
1474: on the parameter $g_{s}N$. Large values of this parameter are
1475: associated with larger regions where dilaton perturbations
1476: associated with the brane are noticeable.
1477:
1478: The solutions associated with supergravity SD-branes induce
1479: dilaton perturbations corresponding to a time-dependent string
1480: coupling, \beq g_{s}(t) = g_{s} e^{\Phi(t)}. \eeq We find that the
1481: time dependence of the dilaton sourced by SD-branes is
1482: qualitatively different when compared to the radial dependence of
1483: the dilaton associated with regular D-branes.\footnote{This is an
1484: other example where features of SD-branes are not simply those
1485: inherited by analytic continuation of D-branes. In fact, a double
1486: analytic continuation ($r\rightarrow i r$ and $t\rightarrow i t$)
1487: of the supergravity D$p$-brane solutions lead to objects with an
1488: imaginary R-R charge. This is unphysical.}
1489:
1490: Figure~\ref{dilaton} shows the time-evolution of the dilaton
1491: function $g_{s}^{-1} e^{\Phi(t)}$ for a SD4-brane with boundary
1492: conditions (\ref{fullbc}). Typically, the function
1493: $g_{s}^{-1}e^\Phi(t)$ decreases from $t=0$ as $t\rightarrow \pm
1494: \infty$. We find that smaller values of $p$ correspond to larger
1495: asymptotic string couplings.
1496:
1497: The dilaton field generated by SD-branes has at least two
1498: interesting properties. First, all solutions are such that the
1499: dilaton stabilizes to a constant asymptotically, \beq
1500: \lim_{t\rightarrow \pm \infty} \Phi(t) = \Phi_{\pm\infty}. \eeq
1501: More generally, the relation between $\Phi_{+\infty}$ and
1502: $\Phi_{-\infty}$ depends on the initial conditions. This last
1503: statement applies to all other bulk fields. Secondly, the
1504: asymptotic value of the dilaton is always smaller than its initial
1505: value at $t=0$, \beq \Phi_{\pm\infty} < \Phi(0). \eeq This implies
1506: that the late/early string coupling ($t\rightarrow \pm \infty$) is
1507: always smaller than the coupling when the tachyon is at the top of
1508: its potential ($t=0$), {\it i.e.}, \beq g_{s} e^{\Phi_{\pm\infty}}
1509: < g_{s} e^{\Phi(0)}. \eeq
1510:
1511: We find that the qualitative features shown on figure
1512: \ref{dilaton} are preserved when the boundary conditions on the
1513: various fields are changed. Nevertheless, we consider the effect
1514: of varying $\Phi(0)$ in some detail. An interesting quantity to
1515: study is the ratio \beq h =
1516: \frac{e^{\Phi_{\pm\infty}}}{e^{\Phi(0)}}, \eeq which gives a
1517: quantitative measure of how much the initial string coupling is
1518: modified asymptotically. The tachyon profile is not altered
1519: significantly when the initial condition on $\Phi(t)$ is varied.
1520: Nevertheless, we observe that that for large values of $\Phi(0)$
1521: (large initial coupling) the tachyon field reaches the bottom of
1522: the potential well faster. A large initial coupling also means
1523: that the bulk fields relax faster to their stable asymptotic
1524: configuration compared to cases where $\Phi(0)$ is
1525: smaller.\footnote{Obviously $\Phi(0)$ can be taken to have
1526: negative values.} The overall effect on the bulk fields is also
1527: enhanced for larger values of $\Phi(0)$. For example, as the
1528: initial coupling is increased the scale factor $a(t)$ stabilizes
1529: to significantly smaller values (see next section). As for the
1530: dilaton field itself, we find that the ratio $h$ is large for
1531: larger values of $\Phi(0)$. For very small values of the initial
1532: string coupling (large negative values of $\Phi(0)$), the ratio
1533: $h$ approaches unity.
1534:
1535: In summary, we find that the parameter $\Phi(0)$, {\it i.e.}, the
1536: parameter determining the string coupling when the tachyon is
1537: close to the top of its potential, strongly determines the {\it
1538: importance} of the unstable brane source effect on the
1539: supergravity bulk modes.
1540:
1541: \FIGURE{\epsfig{file=new_dil.eps}\caption{Time dependence of the
1542: function $e^{\Phi(t)}/g_{s}$ for a SD4-brane with boundary
1543: conditions (\ref{fullbc}).} \label{dilaton}}
1544:
1545: %- - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - +
1546: \subsubsection{Gravitational field}
1547:
1548: We now describe the effect of the unstable brane source on the
1549: time-dependent metric components $a(t)$ and $R(t)$.
1550:
1551: \FIGURE{\epsfig{file=scalefac.eps, width = 10cm}\caption{The scale
1552: factor on the worldvolume of a supergravity SD4-brane with
1553: boundary conditions (\ref{fullbc}).} \label{a}}
1554:
1555: Figure \ref{a} shows the scale factor $a(t)$ on the worldvolume of
1556: a SD4-brane with the boundary conditions (\ref{fullbc}). A general
1557: feature is that away from $t=0$ the scale factor monotonically
1558: decreases from its initial value, $a(0)$, to a stable asymptotic
1559: value, \beq \lim_{t\rightarrow \pm \infty} a(t) = a_{\pm\infty},
1560: \eeq with $a_{+\infty}>a_{-\infty}$. The time it takes for this
1561: scale factor to reach its asymptotic value corresponds roughly to
1562: the time it takes for the tachyon to reach the bottom of its
1563: potential. As pointed out above, for large initial values of the
1564: dilaton, $\Phi(0)$, the asymptotic values of the scale factor,
1565: $a_{\pm\infty}$, become smaller. Correspondingly, when the initial
1566: coupling is weak the effect of the probe on the bulk modes is
1567: small and $a(t)$ stabilizes to a value closer to its initial value
1568: $a(0)$.
1569:
1570: Figure \ref{Rt} shows the behavior of the metric function $R(t)$.
1571: A generic feature of the supergravity SD-brane solutions is that
1572: \beq \lim_{t \rightarrow \pm \infty} R(t) = \pm (t +
1573: \kappa_{\scriptscriptstyle \pm R}). \eeq The constants
1574: $\kappa_{\scriptscriptstyle \pm R}$ are generically larger for
1575: larger values of $p$.
1576:
1577: \FIGURE{\epsfig{file=transscale.eps}\caption{The SD4-brane
1578: transverse scale factor $R(t)$ with boundary conditions
1579: (\ref{fullbc}).} \label{Rt}}
1580:
1581: %- - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - +
1582: \subsubsection{Ramond-Ramond field}
1583:
1584: Figure~\ref{Ct} shows the time dependence of the Ramond-Ramond
1585: form field for a supergravity SD4-brane with boundary conditions
1586: (\ref{fullbc}). Again, the energy stored in this field
1587: (proportional to its time-derivative) goes to zero in
1588: approximately the time it takes for the tachyon to reach the
1589: bottom of its potential. A generic feature of the Ramond-Ramond
1590: field associated with a SD-brane is therefore, \beq
1591: \lim_{t\rightarrow \pm \infty} C(t) = C_{\pm\infty}, \eeq where
1592: $C_{\pm\infty}$ is a constant. Typically we find that these
1593: constants are smaller for larger values of $p$.
1594:
1595: \FIGURE{\epsfig{file=RRnew.eps}\caption{The SD4-brane
1596: Ramond-Ramond field $C(t)$ associated with the boundary conditions
1597: (\ref{fullbc}).} \label{Ct}}
1598:
1599: %- - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - +
1600: \subsubsection{Curvature bounds and asymptotic flatness}
1601:
1602: The supergravity equations of motion are derived from a worldsheet
1603: calculation by requiring that, at a certain order in perturbation
1604: theory, the beta-functions associated with bulk fields vanish.
1605: Typically, there are higher order (in $\alpha'$) curvature
1606: corrections to the beta-functions. These corrections are
1607: negligible only if the curvature involved is small when measured
1608: with respect to the string length, $l_{s} = \sqrt{\alpha'}$. This
1609: is why our solutions can, strictly speaking, be trusted only if
1610: the curvature involved is such that
1611: %
1612: $\left| {\cal R} \right| , \; \left| {\cal R}_{\mu\nu}{\cal
1613: R}^{\mu\nu} \right| , \; \left| {\cal R}_{\mu\nu\rho\lambda}{\cal
1614: R}^{\mu\nu\rho\lambda} \right|$ are small.
1615: %
1616: We verify that this condition is satisfied by studying the
1617: behavior of the time-dependent Ricci scalar of the supergravity
1618: SD-branes,
1619: %
1620: \begin{eqnarray}
1621: \label{curvat} {\cal R}(t) = && 2(p+1)\frac{\ddot{a}}{a} +
1622: 2(8-p)\frac{\ddot{R}}{R}
1623: +2(p+1)(8-p)\frac{\dot{a}}{a}\frac{\dot{R}}{R} \nonumber\\
1624: && +p(p+1)
1625: \left(\frac{\dot{a}^{2}}{a^{2}}+{\frac{k_\parallel}{a^2}}\right)
1626: %
1627: +(8-p)(7-p) \left(
1628: \frac{\dot{R}^{2}}{R^{2}} + \frac{k_{\perp}}{R^{2}} \right),
1629: \end{eqnarray}
1630: %
1631: where $k_{\perp}=-1$ and $k_\parallel=0$ for the cases of interest
1632: here.
1633:
1634: A property of the solutions, which is apparent from studying the
1635: evolution of bulk modes, is that of asymptotic flatness. In fact,
1636: we find \begin{eqnarray} \lim_{t\rightarrow \pm \infty}
1637: ds^{2}_{Sp} = -dt^{2} +
1638: a_{\pm\infty}^{2} d{\vec y}^{2} + (t+\kappa_{\pm R})^{2} dH_{8-p}^{2} \, , \\
1639: \lim_{t\rightarrow \pm \infty} \Phi(t) = \Phi_{\pm\infty},
1640: \;\;\;\; \lim_{t\rightarrow \pm \infty} C(t) = C_{\pm\infty},
1641: \end{eqnarray} where $a_{\pm\infty}$, $\kappa_{\pm R}$, $\Phi_{\pm\infty}$
1642: and $C_{\pm\infty}$ are constants. Both the first- and
1643: second-derivative of the bulk modes vanish asymptotically. The
1644: resulting brane configuration is then clearly flat for
1645: $t\rightarrow \pm \infty$, as it should be.
1646:
1647: An important question to answer at this stage is: {\it Which
1648: quantity in the problem sets an upper bound on the curvature for
1649: SD-branes ?} First, we have found that whatever the curvature is
1650: at $t=0$, its absolute value will never exceed it significantly in
1651: the course of the evolution. This is true only for half S-branes
1652: with boundary conditions corresponding to positive derivatives of
1653: the bulk fields. More generally for full S-branes the acceptable
1654: solutions are those with a combination of the boundary conditions
1655: such that \beq \lim_{t\rightarrow 0} \left| {\cal R}(t) \right| \
1656: {\rm{small}}\ . \eeq These requirements are certainly attainable
1657: within the self-consistent supergravity approximation we are
1658: considering --- along with the specific ansatz we introduced in
1659: order to be able to solve the equations numerically.
1660:
1661: %- - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - +
1662: \subsubsection{The $p=7$ and space-filling SD-branes}
1663:
1664: The case $p=8$ should be special because there is no transverse
1665: space into which ``energy'' can be dissipated. The only bulk
1666: fields involved are then the metric component, $a(t)$, the
1667: dilaton, $\Phi(t)$, and the Ramond-Ramond field, $C(t)$. We find
1668: that the time-derivative of the scale factor, $\dot{a}(t)$,
1669: decreases to zero, as $t\rightarrow \pm \infty$, many orders of
1670: magnitude slower than for $p<7$. The scale factor $a(t)$ also goes
1671: to zero after an infinite time (see section \ref{einstein} for a
1672: physical interpretation) which is to be contrasted with the cases
1673: $p<7$ where $a_{\pm\infty} \neq 0$. One would expect that to be
1674: the source of a curvature singularity at $t=\pm\infty$ but it is
1675: not the case. The Ricci scalar for the space filling SD-brane is
1676: \beq {\cal R}(t) = -72 \left(\frac{\dot{a}}{a}\right)^{2}
1677: -\frac{9}{2} \left(\frac{e^{\Phi}\dot{C}}{a^{9}}\right)^{2} +36
1678: \dot{\Phi} \frac{\dot{a}}{a} + \frac{9}{2}\lambda
1679: e^{\Phi}\Delta^{1/2} V(T). \eeq Curvature singularities are
1680: avoided because all time-derivatives in the problem go to zero
1681: faster than the scale factor as $t\rightarrow \pm \infty$. In
1682: particular, the quantity $e^{\Phi}\dot{C}$ goes to zero faster
1683: than $a(t)$ for large time. Also the string coupling slowly goes
1684: to zero asymptotically ($\Phi_{\pm\infty}\rightarrow -\infty$)!
1685: For the Ramond-Ramond field we find the same behavior as for the
1686: cases $p<7$. The only difference is that the relaxation time of
1687: the bulk modes is many orders of magnitude larger than in the
1688: other cases.
1689:
1690: The functions $a(t)$, $\Phi(t)$ and $C(t)$ associated with the
1691: SD7-brane behave in the same way as the space-filling SD-brane. Of
1692: course in this case there is a transverse space and we find that
1693: the relaxation of $\left| R(t)\right|$ to its asymptotic form
1694: $t+\kappa_{\scriptscriptstyle \pm R}$ takes an infinite amount of
1695: time. In other words, it relaxes to its asymptotic form much
1696: slower than for the cases $p<7$.
1697:
1698: %--------------------------------------------------------------------+
1699: \subsubsection{Einstein frame} \label{einstein}
1700:
1701: Let us end this section with a remark about the physics of our
1702: SD-branes in Einstein frame. This would be the more natural and more
1703: physical frame to use in discussions of potential uses of rolling
1704: tachyons in the context of cosmology.
1705:
1706: The transformation from string frame to Einstein frame involves a
1707: multiplicative factor of $e^{-\Phi/2}$ for $d=10$, which is the
1708: dimension in which we are working here. Now, we have already
1709: commented at length on the behavior of the time-dependent dilaton
1710: field in section \ref{subsec:dildil}. The essential physics there
1711: was that the dilaton is biggest at the top of the potential hill;
1712: in particular, it stabilizes in the infinite past and future at a
1713: {constant} value smaller than the initial condition at the
1714: hilltop. Dilaton derivatives also remain small at all times during
1715: the evolution. Therefore, most of the qualitative features of our
1716: solutions will be preserved upon transformation to Einstein frame;
1717: in particular, all solutions remain completely nonsingular.
1718:
1719: A case that deserves further comments is that of the space filling
1720: SD8-brane discussed earlier. In string frame the metric is written
1721: \beq ds^{2}_{S8} = -dt^{2} + a(t)^{2}d{\vec y}^{2}. \eeq Upon
1722: converting to Einstein frame we get the metric \beq ds_{ES8}^{2} =
1723: -d\tau^{2} + a_{E}(\tau)^{2}d{\vec y}^{2}, \eeq where we have used
1724: a change of coordinate such that $d\tau^{2}=e^{-\Phi/2}dt^{2}$,
1725: and where \beq a_{E}(\tau) = e^{-\Phi(t)/4} a(t). \eeq For the
1726: SD8-brane we found
1727: \begin{eqnarray} \lim_{t\rightarrow \pm\infty} a(t) = 0 \, , \;\;\;\;
1728: \lim_{t\rightarrow \pm\infty} e^{\Phi(t)} = 0.
1729: \end{eqnarray} In string frame this implies that the metric
1730: ``closes~off'' at infinity but in the (more physical) Einstein
1731: frame the converse happens, {\it i.e.}, the limit $\tau\rightarrow
1732: \pm \infty$ corresponds to a constant scale factor, \beq
1733: \lim_{t\rightarrow \pm \infty} a_{E}(\tau) = {\rm const.} \eeq
1734: This is a very sensible result. We mentioned before that, because
1735: of the absence of a transverse space, there appeared to be no
1736: channel into which the energy could go. We therefore find that
1737: asymptotically the energy has gone into inflating the worldvolume
1738: of the SD-brane. In fact, the relaxation time for the
1739: gravitational field is essentially infinite.
1740:
1741: \subsection{Other classes of solutions} \label{othercases} In this section we
1742: present a more general analysis of the family of solutions with
1743: parameters $\{k_{\parallel},k_{\perp}\}$ associated with the metric
1744: ansatz (\ref{genmet}).
1745:
1746: \subsubsection{Comment about the numerical analysis} For
1747: $\lambda=0$ there is clearly no source for the the bulk fields.
1748: The equation of motion for the tachyon is then (\ref{theeom}). In
1749: Appendix \ref{flatach} we found analytic expressions for the
1750: corresponding solutions. They have the large time property \beq
1751: \label{asympbe} \frac{V(T)}{\sqrt{\Delta}} = {\rm const.} \eeq
1752: However, if we solve eq.~(\ref{theeom}) numerically (using the
1753: same techniques as in section~\ref{section:numeric}) we find that
1754: $V(T)/\sqrt{\Delta}$ appears to become constant but, after some
1755: more time has elapsed, its behavior becomes erratic, {\it i.e.},
1756: it starts oscillating with increasingly large amplitudes. Clearly
1757: this is only an artifact of the numerical analysis. The tachyon
1758: evolution is such that \beq \lim_{t\rightarrow \pm \infty} V(T) =
1759: 0, \;\;\; \lim_{t\rightarrow\pm \infty} \sqrt{\Delta} = 0. \eeq
1760: The source of the problem can be traced down to the numerics
1761: having difficulties to evaluate a $\frac{0}{0}$ division at large
1762: times.
1763:
1764: The same thing happens for $\lambda\neq 0$. The quantity that
1765: becomes ambiguous for large time is then \beq \label{quantity}
1766: \frac{e^{\Phi(t)} V(T)}{\sqrt{\Delta}}. \eeq Again, the problem is
1767: associated with the fact that both $e^{\Phi(t)}\, V(T)$ and
1768: $\sqrt{\Delta}$ are zero for large time. It is important to
1769: resolve this ambiguity because the quantity (\ref{quantity})
1770: directly feeds in the equations of motions for the bulk fields. We
1771: are able to show numerically that (at least for the solutions
1772: considered in this work) for large time \beq \frac{e^{\Phi(t)}
1773: V(T)}{\sqrt{\Delta}} = 0. \eeq Similarly to the $\lambda=0$ case,
1774: past some finite time the behavior of the function
1775: (\ref{quantity}) becomes erratic. To obtain numerical solutions
1776: representing the evolution for all times we did the following:
1777: past the time value where (\ref{quantity}) becomes ill-behaved, we
1778: solve the system of differential equations without a source, {\it
1779: i.e.}, for \beq \frac{e^{\Phi(t)} V(T)}{\sqrt{\Delta}} = 0, \;\;\;
1780: e^{\Phi(t)}V(T)\sqrt{\Delta} =0, \eeq and appropriate boundary
1781: conditions. This introduces negligible errors.
1782: %
1783:
1784: \subsubsection{Time-reversal symmetric solutions}\label{trssblar}
1785: We consider the time-reversal symmetric solutions, {\it i.e.},
1786: those associated with bulk fields having vanishing
1787: time-derivatives at $t=0$: \beq
1788: \dot{a}(0)=\dot{R}(0)=\dot{C}(0)=\dot{\Phi}(0)=0. \eeq The
1789: constraint equation (\ref{fconstraint}) at $t=0$ is then \beq
1790: \frac{V(T)}{\sqrt{\Delta}} = \frac{1}{\lambda e^{\Phi}}
1791: \left(p(p+1)\frac{k_{\parallel}}{a^{2}}
1792: +(7-p)(8-p)\frac{k_{\perp}}{R^{2}} \right). \eeq The LHS being
1793: positive-definite, the constraint can be satisfied if and only if
1794: the metric ansatz contains a subspace of positive curvature. These
1795: consist in five categories of solutions, {\it i.e.},
1796: $\{k_{\parallel},k_{\perp}\}=\{0,1\},\{-1,1\},\{1,-1\},\{1,1\},\{1,0\}$.
1797: We have performed a detailed analysis of these solutions for $p=4$
1798: and $\dot{T}(0)\leq 1/10$. Typically, we find that when the
1799: tachyon has reached a point of its evolution where $|V(T)|\simeq
1800: 0$, labelled $t_{c}$, the solutions develop a curvature
1801: singularity. The behavior of the bulk fields is as follows. The
1802: time derivative of the scale factor is such that \beq
1803: \lim_{t\rightarrow t_{c}} \dot{a} = -\infty, \;\;\;
1804: \lim_{t\rightarrow t_{c}} a(t) = 0, \eeq which corresponds to a
1805: big-crunch singularity on the ($p+1$)-dimensional worldvolume. The
1806: cases $\{k_{\parallel}=-1,k_{\perp}=1\}$,
1807: $\{k_{\parallel}=0,k_{\perp}=1\}$ and
1808: $\{k_{\parallel}=1,k_{\perp}=0\}$ are such that \beq
1809: \lim_{t\rightarrow t_{c}} \dot{R}(t) = -\infty, \;\;\;
1810: \lim_{t\rightarrow t_{c}} R(t) =0, \eeq corresponding to the
1811: transverse spherical space collapsing to zero-size in finite time.
1812: For $\{k_{\parallel}=1,k_{\perp}=1\}$ and
1813: $\{k_{\parallel}=1,k_{\perp}=-1\}$, we find \beq
1814: \lim_{t\rightarrow t_{c}} \dot{R}(t) = +\infty, \;\;\;
1815: \lim_{t\rightarrow t_{c}} R(t) = +\infty. \eeq For $k_{\perp}=-1$,
1816: $R(t)$ goes to infinity faster than $t$. All solutions are such
1817: that \beq \lim_{t\rightarrow t_{c}} \dot{\Phi}(t) = -\infty. \eeq
1818: Generically, the large time behavior of the R-R field is
1819: well-behaved, {\it i.e.}, \beq \lim_{t\rightarrow t_{c}}
1820: \dot{C}(t) \approx 0, \;\;\; \lim_{t\rightarrow t_{c}} C(t) =
1821: C_{c},\eeq where $C_{c}$ is finite but typically many orders of
1822: magnitudes larger than the constants $C_{\infty}$ associated with
1823: the regular solutions presented earlier. The case
1824: $\{k_{\parallel}=0,k_{\perp}=-1\}$ is special because then the
1825: time derivative of the R-R field diverges as well at finite time.
1826:
1827: We believe our conclusions to be unaltered for other values of
1828: $p$. We found no evidence that the singularities are resolved when
1829: the time-reversal symmetry is broken.
1830:
1831: \subsubsection{More regular solutions} Another class of candidate
1832: SD-brane solutions we have studied are those with
1833: $k_{\parallel}=0$ and $k_{\perp}=0$. The results we present for
1834: these solutions are generic, {\it i.e.}, they hold for all $p$ and
1835: reasonable boundary conditions ({\it i.e.}, first derivatives not
1836: too large). Firstly, the solutions are always asymmetric around
1837: $t=0$ since it was shown before that, for consistency, at least
1838: one of the bulk fields must have non-vanishing kinetic energy at
1839: $t=0$. The level of asymmetry will be reduced by having smaller
1840: derivatives of the bulk fields at $t=0$. The corresponding
1841: solutions are regular and asymptotically flat. We find that \beq
1842: \lim_{t\rightarrow \pm \infty} \dot{a}(t) = 0, \;\;\;
1843: \lim_{t\rightarrow \pm \infty} a(t) = 0, \eeq and \beq
1844: \lim_{t\rightarrow \pm \infty} \dot{R}(t) = 0, \;\;\;
1845: \lim_{t\rightarrow \pm \infty} R(t) = {\rm const.} \eeq In
1846: Einstein frame both scale factors asymptote to non-zero constants.
1847: For the dilaton we obtain \beq \lim_{t\rightarrow \pm \infty}
1848: \dot{\Phi}(t) = 0, \;\;\; \lim_{t\rightarrow \pm \infty} \Phi(t) =
1849: -\infty, \eeq {\it i.e.}, the string coupling vanishes
1850: asymptotically. Finally the R-R field behaves like \beq
1851: \lim_{t\rightarrow \pm \infty} \dot{C}(t) = 0, \;\;\;
1852: \lim_{t\rightarrow \pm \infty} C(t) = {\rm const.} \eeq We note
1853: that the relaxation time for these solutions is many orders of
1854: magnitude larger than for the $\{k_{\parallel}=0$,
1855: $k_{\perp}=-1\}$ cases presented earlier.
1856:
1857: The two remaining cases are $\{k_{\parallel},k_\perp\}$ equal to
1858: $\{-1,-1\}$ and $\{-1,0\}$. We found evidence that the corresponding
1859: solutions are regular and asymptotically flat.
1860:
1861:
1862: %====================================================================+
1863: \section{Discussion}\label{section:discussion}
1864:
1865: Our primary motivation for this work was the general problem of
1866: seeking mechanisms for resolution of singularities in spacetimes of
1867: interest in string theory. SD-brane supergravity spacetimes presented
1868: in refs.~\cite{gutperle,gutperle2,KMP} create somewhat of a
1869: supergravity {\em emergency} because they are not only singular but
1870: nakedly so.
1871:
1872: An important first step in the resolution program was made in
1873: ref.~\cite{buchel}, where some effects of unstable brane probes in
1874: these backgrounds were considered. In particular, the probe
1875: physics was also sick, and taking this analysis seriously led to
1876: an even more dire assessment of the likelihood of singularity
1877: resolution without resorting to the inclusion of massive string
1878: modes in both the open and closed sectors. In our work, then, we
1879: began by plumbing the depths of the probe approach. We found it to
1880: be generally insufficient for our purposes; one reason is that the
1881: probe approximation takes itself out of its regime of
1882: self-consistency.
1883:
1884: We then launched into an investigation of the physics of the
1885: gravitational fields exerted by SD$p$-branes for general $p$ by
1886: including backreaction. In order to get started on this problem,
1887: we had to make the approximation of considering only the most
1888: relevant open- and closed- string modes, with full gravitational
1889: backreaction taken into account. The equations we derived are
1890: highly nonlinear and couple brane with bulk, so did not lend
1891: themselves to solution analytically. We therefore resorted to
1892: numerical techniques to search for solutions. Because of this
1893: restriction, we had to use an ansatz which smeared the branes in
1894: the transverse space; this allowed us to turn the equations into
1895: ODE's and integrate them numerically. An essential step was to
1896: begin the numerical integration
1897: %
1898: near the {\em top} of the potential hill, and then reconstruct
1899: asymptopia, which we were able to do successfully. Generically, the
1900: solutions are time-reversal {\it asymmetric}. We have shown that the
1901: time-reversal symmetric solutions with the correct R-symmetry are
1902: unavailable. Given our two-stage approximation (lowest modes, and
1903: smeared ansatz), we found it rather satisfying that significant
1904: progress in the resolution program is already found at this level. In
1905: particular, our solutions for rolling tachyons backreacting on
1906: spacetime are {\em completely nonsingular}, and our approximations
1907: satisfy the fundamental property of self-consistency. We find these
1908: conclusions suggestive of resolution of the SD-brane spacetime
1909: singularity emergency.
1910:
1911: It is however hard to know for sure whether our nonsingular results
1912: will survive refinement. Therefore, let us now make some specific
1913: remarks about technical roadblocks we encountered which forced us to
1914: make approximations, their physical consequences, and future outlook.
1915:
1916: Section~\ref{section:prelim} was where we derived the coupled
1917: tachyon-supergravity equations for a general brane distribution,
1918: assuming Chern-Simons terms are turned off in a consistent
1919: truncation. Our resulting equations are on the one hand remarkably
1920: non-robust, and on the other quite robust. What we mean by
1921: non-robustness is this: our ability to obtain nonsingular
1922: evolution depends importantly on the structure of these equations
1923: of motion. Signs are crucial, coefficients are crucial, and so is
1924: the inclusion of Ramond-Ramond fields.
1925: %
1926: In other words, our nonsingular results are highly specific to the
1927: field couplings arising from the low-energy approximation to
1928: string theory. Other ``S-branes'' arising from
1929: ``string-motivated'' actions will probably not possess similarly
1930: nonsingular behavior. The positive type of robustness we refer to
1931: is also a desirable property. What we see manifestly is that the
1932: precise form of the potentials $\{V(T),f(T)\}$ is not important,
1933: apart from the large-$|T|$ behavior which had been derived
1934: elsewhere. The most obvious refinement of our work here will be
1935: to attack the problem of relaxing the requirement of zero NS-NS
1936: $B$-field. Allowing $B^{(2)}$ to be turned on will allow us to
1937: break $ISO(p{+}1)$ on the worldvolume and allow inhomogeneous
1938: tachyonic modes --- the importance of which is discussed in
1939: refs.~\cite{cosmotachyonK,larsen} --- and also to turn on more
1940: components of R-R fields. Inhomogeneities would of course have to
1941: be included in initial conditions, because homogeneous on-shell
1942: tachyons do not couple to non-homogeneous tachyons
1943: \cite{kutasov2}.
1944: %
1945: We have postponed the non-homogeneous problem to the future mainly
1946: because it is messy; our work reported here should be considered a
1947: step in a larger program.
1948:
1949: The other important approximation we made in our work was in
1950: section~\ref{subsection:specific_ansatz}, where we had to smear
1951: the SD-brane sources in the transverse space to facilitate
1952: integrating the equations numerically by turning them into ODE's.
1953: This limits our ability to fully probe the properties of the
1954: system in which we are interested. Here we would also like to
1955: record another physical consequence of this ansatz. Namely, this
1956: restriction has notable, negative, consequences for our ability to
1957: track whether black holes form as intermediate states during the
1958: time evolution of our coupled system including full backreaction.
1959: The issue of black holes was raised in the discussion section of
1960: ref.~\cite{KMP}. The essential point is that a black hole
1961: intermediate state may arise as an alternative to SD-brane
1962: formation and decay, at least with the half-advanced,
1963: half-retarded propagator. The fine-tuned nature of the initial
1964: conditions producing SD-branes highlights a reason why integrating
1965: partial differential equations of motion (including dependence on
1966: transverse coordinates) may be particularly difficult numerically.
1967: Or the obstruction to finding the full solution may yet turn out
1968: to be negotiable. It will also be interesting to think further
1969: about particle/string production.
1970: %
1971:
1972: Let us end with some somewhat speculative remarks. Typically in
1973: the limit $g_{s} N\rightarrow 0$ the open and closed strings
1974: decouple. This is true in our effective lowest-modes analysis
1975: here, but also explicit in other worldsheet-inspired approaches.
1976: There should also exist a limit (in time) to be taken where only
1977: the open strings survive. From the worldsheet definition of a
1978: SD-brane, it is suggestive that the open string degrees of freedom
1979: would combine to form a Euclidean conformal field theory in $p+1$
1980: dimensions. The same appears true when considering the effective
1981: action of massless open string degrees of freedom on an unstable
1982: D-brane \cite{hashimoto}. In both approaches, however, it is not
1983: clear what the role of the tachyon could be. Physically, it is the
1984: source of a process by which energy is siphoned out of the open
1985: string sector and pumped into the closed string sector. So, in a
1986: sense, the decay of a D-brane through tachyon condensation
1987: corresponds to the decrease of a $c$-function-like quantity on the
1988: gauge theory side. Then, we can entertain the idea that
1989: time-evolution on the gravity side should really be regarded as a
1990: renormalization group (RG) flow on the gauge theory side. From
1991: this viewpoint, formation and decay of a SD-brane would be a
1992: process corresponding to first an {\em inverse} RG flow
1993: (integrating in degrees of freedom) followed by regular RG flow
1994: (integrating out degrees of freedom).\footnote{Similar ideas were
1995: explored in the context of the dS/CFT correspondence (see, for
1996: example, refs.~\cite{andycft,us})} This might be related to the
1997: study of open string tachyon condensation using RG flow in the
1998: worldsheet theory \cite{martinec}.
1999:
2000: %--------------------------------------------------------------------+
2001: \vskip0.1\textheight
2002: %--------------------------------------------------------------------+
2003: \section*{Acknowledgements}
2004:
2005: The authors wish to thank Alex Buchel, Lev Kofman, Martin
2006: Kruczenski, Finn Larsen, Juan Maldacena, Alex Maloney, Don Marolf,
2007: Rob Myers, Andy Strominger, and Johannes Walcher for useful
2008: discussions. Finally we thank Dave Winters for proof-reading an
2009: earlier draft of this paper.
2010:
2011: FL was supported in part by NSERC of Canada and FCAR du Qu\'ebec. AWP
2012: thanks the Radcliffe Institute for Advanced Study, and the High-Energy
2013: Theory group of the Harvard University Physics Department, for
2014: hospitality during Fall semester 2002-3 while this work was carried
2015: out. Research of AWP is supported by the Radcliffe Institute, CIAR
2016: and NSERC of Canada, and the Sloan Foundation of the USA.
2017:
2018: %====================================================================+
2019: \newpage
2020: %\vskip0.1\textheight
2021: %--------------------------------------------------------------------+
2022: \appendix
2023: %====================================================================+
2024: \section{Regular KMP SD-brane solutions} \label{regular}
2025:
2026: Among the supergravity solutions found by \cite{KMP}, there exist some
2027: that are regular on the horizon located at $t=\omega$. This is only
2028: realized for the following values of the parameters, \beq \tilde{k} =
2029: 2, \;\;\; H = \frac{4}{7-p}, \;\;\; G=k_{i}=0. \eeq The corresponding
2030: metric is
2031: %
2032: \begin{eqnarray} ds^{2} = && F(t)^{1/2}\alpha(t)^{4/(7-p)}
2033: \left(-dt^{2}+t^{2}dH_{8-p}^{2}\right) \nonumber\\
2034: &&+ F(t)^{-1/2}\left[\sum_{i=2}^{p+1}\left(dx^{i}\right)^{2}
2035: +\left(\frac{\beta(t)}{\alpha(t)}\right)^{2}\left(dx^{1}\right)^{2}
2036: \right].\end{eqnarray}
2037: %
2038: Because these solutions are anisotropic in the worldvolume
2039: directions, it is not clear that they are physically relevant. We
2040: will nevertheless study some interesting properties not considered
2041: in ref.~\cite{KMP}. For example, the region $t=+\omega$ does {\it
2042: not} correspond to an horizon as suggested by the fact that none
2043: of the metric components either vanish or diverge there. For $p$
2044: {\it odd} the solutions are time-reversal symmetric so the region
2045: $t=-\omega$ is also not an horizon.
2046:
2047: %--------------------------------------------------------------------+
2048: \subsection{The region close to the origin}
2049:
2050: For the anisotropic solutions there is no curvature singularity at
2051: $t=\omega$. It is therefore interesting to consider the behavior
2052: of the metric components and curvature invariants close to the
2053: potentially problematic region around the origin, $t=0$. We
2054: evaluated the curvature invariants: ${\cal R}$, ${\cal
2055: R}_{\mu\nu}{\cal R}^{\mu\nu}$, and ${\cal
2056: R}_{\mu\nu\rho\lambda}{\cal R}^{\mu\nu\rho\lambda}$. They
2057: identically vanish for $t=0$. The metric tensor there appears
2058: suspicious (for example, the component $g_{tt}$ diverges) but we
2059: find \beq \label{limit} \lim_{t\rightarrow 0} ds^{2} = - d\tau^{2}
2060: + \tau^{2}dH_{8-p}^{2} + \sum_{i=1}^{p+1}\left(dx^{i}\right)^{2},
2061: \eeq where $\tau=\omega^{2}/t$. The expression (\ref{limit}) is
2062: simply flat space with part of it written in Milne coordinates.
2063: Not surprisingly, we find \beq \lim_{t\rightarrow 0} \dot{\Phi}(t)
2064: = 0, \;\;\;\; \lim_{t\rightarrow 0} \dot{C}(t) = 0, \eeq which
2065: implies that all stress-energy components vanish in the the region
2066: close to the origin.
2067:
2068: %--------------------------------------------------------------------+
2069: \subsection{Horizon physics}
2070:
2071: We demonstrate that, contrary to previous claims, many of the
2072: anisotropic solutions are actually regular in the {\em full}
2073: range: $-\infty<t<+\infty$. The anisotropic solutions were
2074: already shown to be non-singular at $t=\omega$ and $t=0$. We now
2075: investigate the region $t=-\omega$ further. Let us introduce the
2076: change of coordinates \beq T = \left( 1 +
2077: \left(\frac{\omega}{t}\right)^{7-p} \right)^{2/(7-p)}t \eeq in
2078: order to make comparison with the results of ref.~\cite{KMP}
2079: easier. For $p$ even, $T=0$ corresponds to $t=-\omega$ while for
2080: $p$ odd we have $T=-2\omega$ when $t=-\omega$. A comment in
2081: ref.~\cite{KMP} is that $T=0$ corresponds to a (non-naked)
2082: curvature singularity. Actually, this is not always the case! For
2083: example, we considered the case $p=1$ and computed the associated
2084: curvature invariants at all times. Figure \ref{fig1} shows the
2085: evolution of the Ricci scalar for the solution with $\omega=1$ and
2086: $\theta=\pi/4$. Clearly, the $p=1$ solution is symmetric under
2087: time-reversal and therefore has no curvature singularity. It also
2088: does not possess any horizon, a feature common to the regular
2089: solutions found in this paper. The qualitative behavior of all
2090: curvature invariants is similar to the Ricci scalar and is quite
2091: generic, {\it i.e.}, it is unchanged for all {\it odd} values of
2092: $p$, $\theta$ and $\omega$. For $p=1$ we obtain \beq
2093: \lim_{t\rightarrow \pm \omega} {\cal R} = -\frac{3
2094: 2^{1/3}}{\omega^{2}}\frac{19\cos^{4}\theta -26\cos^{2}\theta +
2095: 7}{\sin^{5}\theta}. \eeq This is finite except for $\theta=0$ in
2096: which case the Ricci scalar diverges like $R\sim
2097: 1/(t-\omega)^{3}$. We also found expressions for two other
2098: curvature invariants (for $p=1$), \beq \lim_{t\rightarrow \pm
2099: \omega} {\cal R}_{\mu\nu}{\cal R}^{\mu\nu} =\frac{9
2100: 2^{3/4}}{2\omega^{4}} \frac{-272 \cos^{6}\theta + 14 + 255
2101: \cos^{4}\theta + 101\cos^{8}\theta
2102: -98\cos^{2}\theta}{\sin^{10}\theta}, \eeq \beq \lim_{t\rightarrow
2103: \pm \omega} {\cal R}_{\mu\nu\rho\sigma}{\cal R}^{\mu\nu\rho\sigma}
2104: = \frac{3 2^{2/3}}{4 \omega^{4}}\frac{-112\cos \theta + 28 +892
2105: \cos^{8}\theta -1840 \cos^{6}\theta +1032 \cos^{8}\theta
2106: }{\sin^{10}\theta}. \eeq We found expressions with similar
2107: qualitative behavior for other values of $p$ {\it odd}.
2108:
2109: For $p$ {\it even} the solutions are not time-reversal symmetric.
2110: As pointed out previously, curvature invariants are finite at
2111: $t=\omega$ but there is a curvature singularity at $t=-\omega$.
2112: These are the spacelike curvature singularities (protected by an
2113: horizon at $t=0$) described in ref.~\cite{KMP}.
2114:
2115: \FIGURE{\epsfig{file=ricci1.eps}\caption{This figure illustrates
2116: the Ricci scalar for the $p=1$ anisotropic SD-brane solution with
2117: $\omega=1$ and $\theta=\pi/4$. The other curvature invariants
2118: behave similarly.} \label{fig1}}
2119:
2120: %--------------------------------------------------------------------+
2121: \subsection{Unstable brane probe analysis}
2122:
2123: As mentioned in section \ref{subsection:buchel}, the motivation
2124: behind considering an unstable brane probe in a background with
2125: singularity problems is to ask if the singular background could
2126: actually be built.
2127:
2128: The calculations and results of ref.~\cite{buchel} were summarized
2129: in section \ref{subsection:buchel}. In this appendix we generalize
2130: this calculation by probing the $d=10$ {\em anisotropic}
2131: backgrounds presented above, since these are the only ones which
2132: are either non-singular or have singularities (at $t=-\omega$)
2133: shielded by a horizon ($t=0$). We felt this generalization to be
2134: necessary because ref.~\cite{buchel} did not, for example, address
2135: the issue as to how the inclusion of the dilaton might affect the
2136: brane probe calculation. The unstable brane action is the obvious
2137: generalization eq.~(\ref{sourcebuc}) of the case studied in
2138: ref.~\cite{buchel}.
2139:
2140: We investigate whether or not an unstable brane probe is a
2141: well-defined object in the vicinity of the region $t=\omega$. In
2142: Einstein frame, the energy density for the probe propagating in
2143: the anisotropic backgrounds is \beq \rho_{probe} =\frac{N
2144: \mu_{p+1}}{g_{s}} f(\Phi) \frac{V(T)}{\Delta^{1/2}}, \eeq while
2145: the pressure corresponds to \beq p_{probe} =\frac{N
2146: \mu_{p+1}}{g_{s}} f(\Phi) V(T) \Delta^{1/2}. \eeq The dilaton
2147: function $f(\Phi)$ was picked up during the transformation from
2148: the string frame to the Einstein frame. It plays no role in the
2149: upcoming analysis because the dilaton is well-behaved, \beq
2150: \lim_{t\rightarrow \omega} f(\Phi)= {\rm const.} \eeq
2151:
2152: As we saw in section \ref{subsection:buchel}, whenever the probe
2153: analysis goes wrong, it signals a pathology for the gravitational
2154: background. As we saw, there are at least two ways the probe
2155: analysis can go wrong: it may induce an infinite energy or
2156: pressure density ($\rho_{probe}, p_{probe} \rightarrow \pm
2157: \infty$), or, there might not exist any reasonable solutions for
2158: $T(t)$.
2159:
2160: For $t\simeq \omega$, the dominant contribution to the equation of
2161: motion for the tachyon is \beq \Delta^{2} - \Delta + \frac{2}{9}
2162: (t-\omega) \dot{\Delta} = 0. \eeq This is solved for \beq
2163: \Delta(t) = \frac{(t-\omega)^{9/2}}{(t-\omega)^{9/2}-g},\eeq where
2164: $g$ is a constant of integration. The solution $\Delta=0$ ($g \neq
2165: 0$) clearly corresponds to the brane probe inducing a curvature
2166: singularity on the horizon. The only physical solution is the one
2167: for which $g=0$ which corresponds to $\Delta=1$. For the
2168: anisotropic backgrounds considered here the metric component
2169: $g_{tt}$ neither vanishes nor blows up at $t=\omega$. The solution
2170: $\Delta=1$ therefore corresponds to a tachyon field for which the
2171: time-derivative vanishes ($\dot{T}=0$) at $t=\omega$. It therefore
2172: appears that there are solutions for the probe evolution that
2173: avoids the pathologies described earlier. This is no surprise
2174: since for these anisotropic backgrounds the region $t=\omega$ is
2175: not an horizon in the technical sense of the term. We repeated the
2176: calculation around the regions $t=0$ and $t=-\omega$. We find that
2177: for both $p$ {\it odd} and {\it even} the unstable brane probe is
2178: not well-behaved at $t=0$, {\it i.e.}, it induces a curvature
2179: singularity there.
2180:
2181: %====================================================================+
2182: \section{Tachyon in flat space}\label{flatach}
2183:
2184: We consider solutions to the equation of motion for an open string
2185: tachyon when the massless closed string modes are decoupled. The
2186: relevant equation of motion is \beq \label{eomt_two} \ddot{T} +
2187: (1-\dot{T}^{2})\frac{\partial \ln V(T)}{\partial T} = 0. \eeq
2188:
2189: \subsection{General solution} Eq.~(\ref{eomt_two}) is a second
2190: order differential equation with a general solution of the form
2191: \beq T(t) = \int dt \; \frac{1 + a^{2}\, V^{2}(t)}{1-a^{2}\,
2192: V^{2}(t)} + b, \eeq where $a$ and $b$ are constants of
2193: integration. To integrate this equation one needs the function
2194: $V(t)$ which would imply that we already know the solution $T(t)$.
2195: Open string field theory has taught us that for $t=t_{c}$ large we
2196: have \beq \lim_{t \rightarrow t_{c}} V(t) << 1, \eeq which implies
2197: that \beq T(t) \simeq \int_{t_{c}}^{t} dt \; (1 + 2 a^{2} V(t)) +
2198: b. \eeq Therefore at large time the tachyon behaves like \beq T(t)
2199: = t + (b-t_{c}) + 2a^{2} \int_{t_{c}}^{t} dt \; V^{2}(t). \eeq
2200: Using the string field theory result \beq \lim_{T\rightarrow
2201: +\infty} V(T) = e^{-T/\sqrt{2}}, \eeq and taking $T(t)\simeq t$
2202: leads to the large time formula \beq T(t) \simeq t -a^{2}\sqrt{2}
2203: e^{-\sqrt{2}t}, \eeq where we have fixed the integration constants
2204: by imposing \beq a^{2}\sqrt{2} e^{-\sqrt{2}t_{c}} + b - t_{c} =
2205: 0.\eeq
2206:
2207: \subsection{Particular solutions} We consider solutions to the
2208: equation of motion (\ref{eomt_two}) with the potential \beq V(T) =
2209: \frac{1}{\cosh \left(T/\sqrt{2}\right)} \, . \eeq The tachyon
2210: equation of motion becomes \beq \label{theeom} \ddot{T} +
2211: \frac{1}{\sqrt{2}}(1-\dot{T}^{2}) \tanh \left(T/\sqrt{2}\right),
2212: \eeq which has a solution of the form \beq T(t) = -\sqrt{2}\; {\rm
2213: arc\, sinh}\; \left( \frac{\sqrt{2}}{2} \left[ c_{1}
2214: e^{t/\sqrt{2}} - c_{2} e^{-t/\sqrt{2}} \right] \right), \eeq where
2215: $c_{1}$ and $c_{2}$ are constants of integration. We usually
2216: specify boundary conditions at $t=0$, \beq T(0)=-\sqrt{2} \;{\rm
2217: arc\, sinh} \; \left(\frac{\sqrt{2}(c_{1}-c_{2})}{2}\right), \eeq
2218: \beq \dot{T}(0)=-\sqrt{2}
2219: \frac{c_{1}+c_{2}}{\left(4+2(c_{1}-c_{2})^{2}\right)^{1/2}}. \eeq
2220: The family of solutions characterized by $T(0)=0$ ($c_{1}=c_{2}$)
2221: corresponds to all possible tachyon velocities at $t=0$:
2222: $\dot{T}(0)=-\sqrt{2} c_{1}$. An other class of solutions are
2223: those for which $\dot{T}(0)=0$ ($c_{1}=-c_{2}$). Those correspond
2224: to allowing the tachyon to begin its evolution with
2225: $T(0)=-\sqrt{2}\;{\rm arcsinh} \; \sqrt{2}c_{1}$.
2226:
2227: The solution we presented are referred to as {\it tachyon matter}.
2228: The stress-energy components (which are independent of the number
2229: of dimensions in the theory) correspond to a conserved energy
2230: density ($\rho \sim V(T)/\sqrt{\Delta} = {\rm constant}$) and a
2231: pressure ($p\sim -V(T)\sqrt{\Delta}$) that vanishes as
2232: $t\rightarrow +\infty$ \cite{sen2}.
2233:
2234: %====================================================================+
2235: \newpage
2236:
2237: \begin{thebibliography}{99}
2238:
2239: \bibitem{gutperle} M.~Gutperle and A.~Strominger, ``Spacelike
2240: branes,'' JHEP {\bf 0204}, 018 (2002) [arXiv:hep-th/0202210].
2241: %%CITATION = HEP-TH 0202210;%%
2242:
2243: \bibitem{dscft}
2244: A.~Strominger, ``The dS/CFT correspondence,'' JHEP {\bf 0110}, 034
2245: (2001) [arXiv:hep-th/0106113]; %%CITATION = HEP-TH 0106113;%%
2246: E.~Witten, ``Quantum gravity in de Sitter space,''
2247: arXiv:hep-th/0106109. %%CITATION = HEP-TH 0106109;%%
2248:
2249: \bibitem{tale}
2250: F.~Leblond, D.~Marolf and R.~C.~Myers, ``Tall tales from de Sitter
2251: space II: Field theory dualities,'' JHEP {\bf 0301}, 003 (2003)
2252: [arXiv:hep-th/0211025].
2253: %%CITATION = HEP-TH 0211025;%%
2254:
2255: \bibitem{gutperle2}
2256: C.~M.~Chen, D.~V.~Gal'tsov and M.~Gutperle, ``S-brane solutions in
2257: supergravity theories,'' Phys.\ Rev.\ D {\bf 66}, 024043 (2002)
2258: [arXiv:hep-th/0204071].
2259: %%CITATION = HEP-TH 0204071;%%
2260:
2261: \bibitem{KMP}
2262: M.~Kruczenski, R.~C.~Myers and A.~W.~Peet, ``Supergravity
2263: S-branes,'' JHEP {\bf 0205}, 039 (2002) [arXiv:hep-th/0204144].
2264: %%CITATION = HEP-TH 0204144;%%
2265:
2266: \bibitem{strominger}
2267: A.~Strominger, ``Open string creation by S-branes,''
2268: arXiv:hep-th/0209090.
2269: %%CITATION = HEP-TH 0209090;%%
2270:
2271: \bibitem{buchel}
2272: A.~Buchel, P.~Langfelder and J.~Walcher, ``Does the tachyon
2273: matter?,'' Annals Phys.\ {\bf 302}, 78 (2002)
2274: [arXiv:hep-th/0207235]; A.~Buchel and J.~Walcher, ``The tachyon
2275: does matter,'' arXiv:hep-th/0212150.
2276: %%CITATION = HEP-TH 0212150;%%
2277: %%CITATION = HEP-TH 0207235;%%
2278:
2279: \bibitem{roy}
2280: S.~Roy, ``On supergravity solutions of space-like Dp-branes,''
2281: JHEP {\bf 0208}, 025 (2002) [arXiv:hep-th/0205198].
2282: %%CITATION = HEP-TH 0205198;%%
2283:
2284: \bibitem{sbranes}
2285: B.~McInnes, ``dS/CFT, censorship, instability of hyperbolic
2286: horizons, and spacelike branes,'' arXiv:hep-th/0205103;
2287: %%CITATION = HEP-TH 0205103;%%
2288: N.~S.~Deger and A.~Kaya, ``Intersecting S-brane solutions of D =
2289: 11 supergravity,'' JHEP {\bf 0207}, 038 (2002)
2290: [arXiv:hep-th/0206057];
2291: %%CITATION = HEP-TH 0206057;%%
2292: J.~E.~Wang, ``Spacelike and time dependent branes from DBI,'' JHEP
2293: {\bf 0210}, 037 (2002) [arXiv:hep-th/0207089];
2294: %%CITATION = HEP-TH 0207089;%%
2295: C.~P.~Burgess, F.~Quevedo, S.~J.~Rey, G.~Tasinato and I.~Zavala,
2296: ``Cosmological spacetimes from negative tension brane
2297: backgrounds,'' JHEP {\bf 0210}, 028 (2002) [arXiv:hep-th/0207104];
2298: %%CITATION = HEP-TH 0207104;%%
2299: V.~D.~Ivashchuk, ``Composite S-brane solutions related to
2300: Toda-type systems,'' Class.\ Quant.\ Grav.\ {\bf 20}, 261 (2003)
2301: [arXiv:hep-th/0208101];
2302: %%CITATION = HEP-TH 0208101;%%
2303: F.~Quevedo, G.~Tasinato and I.~Zavala, ``S-branes, negative
2304: tension branes and cosmology,'' arXiv:hep-th/0211031;
2305: %%CITATION = HEP-TH 0211031;%%
2306: N.~Ohta, ``Intersection rules for S-branes,''
2307: arXiv:hep-th/0301095;
2308: %%CITATION = HEP-TH 0301095;%%
2309: C.~P.~Burgess, P.~Martineau, F.~Quevedo, G.~Tasinato and I.~Zavala
2310: C., ``Instabilities and particle production in S-brane
2311: geometries,'' arXiv:hep-th/0301122.
2312: %%CITATION = HEP-TH 0301122;%%
2313:
2314: \bibitem{andylastweek}
2315: A.~Maloney, A.~Strominger and X.~Yin, ``S-brane thermodynamics,''
2316: arXiv:hep-th/0302146.
2317: %%CITATION = HEP-TH 0302146;%%
2318:
2319: \bibitem{strominger3}
2320: M.~Gutperle and A.~Strominger, ``Timelike boundary Liouville
2321: theory,'' arXiv:hep-th/0301038.
2322: %%CITATION = HEP-TH 0301038;%%
2323:
2324: \bibitem{vacuum}
2325: A.~Sen, ``Fundamental strings in open string theory at the
2326: tachyonic vacuum,'' J.\ Math.\ Phys.\ {\bf 42}, 2844 (2001)
2327: [arXiv:hep-th/0010240];
2328: %%CITATION = HEP-TH 0010240;%%
2329: L.~Rastelli, A.~Sen and B.~Zwiebach, ``String field theory around
2330: the tachyon vacuum,'' Adv.\ Theor.\ Math.\ Phys.\ {\bf 5}, 353
2331: (2002) [arXiv:hep-th/0012251].
2332: %%CITATION = HEP-TH 0012251;%%
2333:
2334: \bibitem{OS}
2335: T.~Okuda and S.~Sugimoto, ``Coupling of rolling tachyon to closed
2336: strings,'' Nucl.\ Phys.\ B {\bf 647}, 101 (2002)
2337: [arXiv:hep-th/0208196].
2338: %%CITATION = HEP-TH 0208196;%%
2339:
2340: \bibitem{maldacena}
2341: N.~Lambert, H.~Liu and J.~Maldacena, ``Closed strings from
2342: decaying D-branes,'' arXiv:hep-th/0303139.
2343: %%CITATION = HEP-TH 0303139;%%
2344:
2345: \bibitem{sen2}
2346: A.~Sen, ``Rolling tachyon,'' JHEP {\bf 0204}, 048 (2002)
2347: [arXiv:hep-th/0203211];
2348: %%CITATION = HEP-TH 0203211;%%
2349: A.~Sen, ``Tachyon matter,'' JHEP {\bf 0207}, 065 (2002)
2350: [arXiv:hep-th/0203265];
2351: %%CITATION = HEP-TH 0203265;%%
2352: A.~Sen, ``Field theory of tachyon matter,'' Mod.\ Phys.\ Lett.\ A
2353: {\bf 17}, 1797 (2002) [arXiv:hep-th/0204143].
2354: %%CITATION = HEP-TH 0204143;%%
2355:
2356: \bibitem{gibbons}
2357: G.~W.~Gibbons, ``Cosmological evolution of the rolling tachyon,''
2358: Phys.\ Lett.\ B {\bf 537}, 1 (2002) [arXiv:hep-th/0204008].
2359: %%CITATION = HEP-TH 0204008;%%
2360:
2361: \bibitem{cosmotachyonK}
2362: %%CITATION = HEP-TH 0204150;%%
2363: A.~V.~Frolov, L.~Kofman and A.~A.~Starobinsky, ``Prospects and
2364: problems of tachyon matter cosmology,'' Phys.\ Lett.\ B {\bf 545},
2365: 8 (2002) [arXiv:hep-th/0204187];
2366: %%CITATION = HEP-TH 0205003;%%
2367: L.~Kofman and A.~Linde, ``Problems with tachyon inflation,'' JHEP
2368: {\bf 0207}, 004 (2002) [arXiv:hep-th/0205121];
2369: %%CITATION = HEP-TH 0207156;%%
2370: G.~N.~Felder, L.~Kofman and A.~Starobinsky, ``Caustics in tachyon
2371: matter and other Born-Infeld scalars,'' JHEP {\bf 0209}, 026
2372: (2002) [arXiv:hep-th/0208019].
2373:
2374: \bibitem{cosmotachyon}
2375: M.~Fairbairn and M.~H.~Tytgat, ``Inflation from a tachyon
2376: fluid?,'' Phys.\ Lett.\ B {\bf 546}, 1 (2002)
2377: [arXiv:hep-th/0204070];
2378: %%CITATION = HEP-TH 0204070;%%
2379: S.~Mukohyama, ``Brane cosmology driven by the rolling tachyon,''
2380: Phys.\ Rev.\ D {\bf 66}, 024009 (2002) [arXiv:hep-th/0204084];
2381: %%CITATION = HEP-TH 0204084;%%
2382: A.~Feinstein, ``Power-law inflation from the rolling tachyon,''
2383: Phys.\ Rev.\ D {\bf 66}, 063511 (2002) [arXiv:hep-th/0204140];
2384: %%CITATION = HEP-TH 0204140;%%
2385: T.~Padmanabhan, ``Accelerated expansion of the universe driven by
2386: tachyonic matter,'' Phys.\ Rev.\ D {\bf 66}, 021301 (2002)
2387: [arXiv:hep-th/0204150];
2388: %%CITATION = HEP-TH 0204187;%%
2389: D.~Choudhury, D.~Ghoshal, D.~P.~Jatkar and S.~Panda, ``On the
2390: cosmological relevance of the tachyon,'' Phys.\ Lett.\ B {\bf
2391: 544}, 231 (2002) [arXiv:hep-th/0204204];
2392: %%CITATION = HEP-TH 0204204;%%
2393: X.~z.~Li, J.~g.~Hao and D.~j.~Liu, ``Can quintessence be the
2394: rolling tachyon?,'' Chin.\ Phys.\ Lett.\ {\bf 19}, 1584 (2002)
2395: [arXiv:hep-th/0204252];
2396: %%CITATION = HEP-TH 0204252;%%
2397: G.~Shiu and I.~Wasserman, ``Cosmological constraints on tachyon
2398: matter,'' Phys.\ Lett.\ B {\bf 541}, 6 (2002)
2399: [arXiv:hep-th/0205003];
2400: %%CITATION = HEP-TH 0205121;%%
2401: H.~B.~Benaoum, ``Accelerated universe from modified Chaplygin gas
2402: and tachyonic fluid,'' arXiv:hep-th/0205140;
2403: %%CITATION = HEP-TH 0205140;%%
2404: M.~Sami, ``Implementing power law inflation with rolling tachyon
2405: on the brane,'' arXiv:hep-th/0205146;
2406: %%CITATION = HEP-TH 0205146;%%
2407: M.~Sami, P.~Chingangbam and T.~Qureshi, ``Aspects of tachyonic
2408: inflation with exponential potential,'' Phys.\ Rev.\ D {\bf 66},
2409: 043530 (2002) [arXiv:hep-th/0205179];
2410: %%CITATION = HEP-TH 0205179;%%
2411: J.~c.~Hwang and H.~Noh, ``Cosmological perturbations in a
2412: generalized gravity including tachyonic condensation,'' Phys.\
2413: Rev.\ D {\bf 66}, 084009 (2002) [arXiv:hep-th/0206100];
2414: %%CITATION = HEP-TH 0206100;%%
2415: T.~Mehen and B.~Wecht, ``Gauge fields and scalars in rolling
2416: tachyon backgrounds,'' arXiv:hep-th/0206212;
2417: %%CITATION = HEP-TH 0206212;%%
2418: K.~Ohta and T.~Yokono, ``Gravitational approach to tachyon
2419: matter,'' Phys.\ Rev.\ D {\bf 66}, 125009 (2002)
2420: [arXiv:hep-th/0207004];
2421: %%CITATION = HEP-TH 0207004;%%
2422: Y.~S.~Piao, R.~G.~Cai, X.~m.~Zhang and Y.~Z.~Zhang, ``Assisted
2423: tachyonic inflation,'' Phys.\ Rev.\ D {\bf 66}, 121301 (2002)
2424: [arXiv:hep-ph/0207143];
2425: %%CITATION = HEP-PH 0207143;%%
2426: G.~Shiu, S.~H.~Tye and I.~Wasserman, ``Rolling tachyon in brane
2427: world cosmology from superstring field theory,''
2428: arXiv:hep-th/0207119;
2429: %%CITATION = HEP-TH 0207119;%%
2430: X.~z.~Li, D.~j.~Liu and J.~g.~Hao, ``On the tachyon inflation,''
2431: arXiv:hep-th/0207146;
2432: %%CITATION = HEP-TH 0207146;%%
2433: J.~M.~Cline, H.~Firouzjahi and P.~Martineau, ``Reheating from
2434: tachyon condensation,'' JHEP {\bf 0211}, 041 (2002)
2435: [arXiv:hep-th/0207156];
2436: %%CITATION = HEP-TH 0208019;%%
2437: B.~Wang, E.~Abdalla and R.~K.~Su, ``Dynamics and holographic
2438: discreteness of tachyonic inflation,'' arXiv:hep-th/0208023;
2439: %%CITATION = HEP-TH 0208023;%%
2440: S.~Mukohyama, ``Inhomogeneous tachyon decay, light-cone structure
2441: and D-brane network problem in tachyon cosmology,'' Phys.\ Rev.\
2442: D {\bf 66}, 123512 (2002) [arXiv:hep-th/0208094];
2443: %%CITATION = HEP-TH 0208094;%%
2444: M.~C.~Bento, O.~Bertolami and A.~A.~Sen, ``Tachyonic inflation in
2445: the braneworld scenario,'' arXiv:hep-th/0208124;
2446: %%CITATION = HEP-TH 0208124;%%
2447: J.~g.~Hao and X.~z.~Li, ``Reconstructing the equation of state of
2448: tachyon,'' Phys.\ Rev.\ D {\bf 66}, 087301 (2002)
2449: [arXiv:hep-th/0209041];
2450: %%CITATION = HEP-TH 0209041;%%
2451: C.~j.~Kim, H.~B.~Kim and Y.~b.~Kim, ``Rolling tachyons in string
2452: cosmology,'' Phys.\ Lett.\ B {\bf 552}, 111 (2003)
2453: [arXiv:hep-th/0210101];
2454: %%CITATION = HEP-TH 0210101;%%
2455: F.~Quevedo, ``Lectures on string / brane cosmology,'' Class.\
2456: Quant.\ Grav.\ {\bf 19}, 5721 (2002) [arXiv:hep-th/0210292];
2457: %%CITATION = HEP-TH 0210292;%%
2458: G.~Shiu, ``Tachyon dynamics and brane cosmology,''
2459: arXiv:hep-th/0210313;
2460: %%CITATION = HEP-TH 0210313;%%
2461: J.~S.~Bagla, H.~K.~Jassal and T.~Padmanabhan, ``Cosmology with
2462: tachyon field as dark energy,'' arXiv:astro-ph/0212198;
2463: %%CITATION = ASTRO-PH 0212198;%%
2464: Y.~S.~Piao, Q.~G.~Huang, X.~m.~Zhang and Y.~Z.~Zhang,
2465: ``Non-minimally coupled tachyon and inflation,''
2466: arXiv:hep-ph/0212219;
2467: %%CITATION = HEP-PH 0212219;%%
2468: X.~z.~Li and X.~h.~Zhai, ``The tachyon inflationary models with
2469: exact mode functions,'' arXiv:hep-ph/0301063;
2470: %%CITATION = HEP-PH 0301063;%%
2471: G.~W.~Gibbons, ``Thoughts on tachyon cosmology,''
2472: arXiv:hep-th/0301117;
2473: %%CITATION = HEP-TH 0301117;%%
2474: M.~Sami, P.~Chingangbam and T.~Qureshi, ``Cosmological aspects of
2475: rolling tachyon,'' arXiv:hep-th/0301140;
2476: %%CITATION = HEP-TH 0301140;%%
2477: C.~j.~Kim, H.~B.~Kim, Y.~b.~Kim and O.~K.~Kwon, ``Cosmology of
2478: rolling tachyon,'' arXiv:hep-th/0301142.
2479: %%CITATION = HEP-TH 0301142;%%
2480:
2481: \bibitem{emparan}
2482: R.~Emparan and H.~S.~Reall, ``A rotating black ring in five
2483: dimensions,'' Phys.\ Rev.\ Lett.\ {\bf 88}, 101101 (2002)
2484: [arXiv:hep-th/0110260].
2485: %%CITATION = HEP-TH 0110260;%%
2486:
2487: \bibitem{gibbonshair} G.~W.~Gibbons, D.~Ida and T.~Shiromizu,
2488: ``Uniqueness and non-uniqueness of static black holes in higher
2489: dimensions,'' Phys.\ Rev.\ Lett.\ {\bf 89}, 041101 (2002)
2490: [arXiv:hep-th/0206049];
2491: %%CITATION = HEP-TH 0206049;%%
2492: %G.~W.~Gibbons, D.~Ida and T.~Shiromizu,
2493: ``Uniqueness of (dilatonic) charged black holes and black p-branes in
2494: higher dimensions,'' Phys.\ Rev.\ D {\bf 66}, 044010 (2002)
2495: [arXiv:hep-th/0206136].
2496: %%CITATION = HEP-TH 0206136;%%
2497:
2498: \bibitem{BMPV}
2499: J.~C.~Breckenridge, R.~C.~Myers, A.~W.~Peet and C.~Vafa,
2500: ``D-branes and spinning black holes,''
2501: Phys.\ Lett.\ B {\bf 391}, 93 (1997)
2502: [arXiv:hep-th/9602065].
2503: %%CITATION = HEP-TH 9602065;%%
2504:
2505: \bibitem{reallbmpv}
2506: H.~S.~Reall,
2507: ``Higher dimensional black holes and supersymmetry,''
2508: arXiv:hep-th/0211290.
2509: %%CITATION = HEP-TH 0211290;%%
2510:
2511: \bibitem{IMSY} N.~Itzhaki, J.~M.~Maldacena, J.~Sonnenschein and
2512: S.~Yankielowicz, ``Supergravity and the large N limit of theories with
2513: sixteen supercharges,'' Phys.\ Rev.\ D {\bf 58}, 046004 (1998)
2514: [arXiv:hep-th/9802042].
2515: %%CITATION = HEP-TH 9802042;%%
2516:
2517: \bibitem{awpconfproc} A.~W.~Peet, ``More on singularity resolution,''
2518: published in the proceedings of {\em Strings 2001} (World Scientific,
2519: 2002) [arXiv:hep-th/0106148].
2520: %%CITATION = HEP-TH 0106148;%%
2521:
2522: \bibitem{add1}
2523: A.~Sen, ``Supersymmetric world-volume action for non-BPS
2524: D-branes,'' JHEP {\bf 9910}, 008 (1999) [arXiv:hep-th/9909062].
2525: %%CITATION = HEP-TH 9909062;%%
2526:
2527: \bibitem{add2}
2528: M.~R.~Garousi, ``On-shell S-matrix and tachyonic effective
2529: actions,'' Nucl.\ Phys.\ B {\bf 647}, 117 (2002)
2530: [arXiv:hep-th/0209068].
2531: %%CITATION = HEP-TH 0209068;%%
2532:
2533: \bibitem{add3}
2534: E.~A.~Bergshoeff, M.~de Roo, T.~C.~de Wit, E.~Eyras and S.~Panda,
2535: ``T-duality and actions for non-BPS D-branes,'' JHEP {\bf 0005},
2536: 009 (2000) [arXiv:hep-th/0003221].
2537: %%CITATION = HEP-TH 0003221;%%
2538:
2539: \bibitem{add4}
2540: J.~Kluson, ``Proposal for non-BPS D-brane action,'' Phys.\ Rev.\ D
2541: {\bf 62}, 126003 (2000) [arXiv:hep-th/0004106].
2542: %%CITATION = HEP-TH 0004106;%%
2543:
2544: \bibitem{sen1}
2545: A.~Sen, ``Time and tachyon,'' arXiv:hep-th/0209122.
2546: %%CITATION = HEP-TH 0209122;%%
2547:
2548: \bibitem{enhanconjpp} C.~V.~Johnson, A.~W.~Peet and J.~Polchinski,
2549: ``Gauge theory and the excision of repulson singularities,''
2550: Phys.\ Rev.\ D {\bf 61}, 086001 (2000) [arXiv:hep-th/9911161].
2551: %%CITATION = HEP-TH 9911161;%%
2552:
2553: \bibitem{polchinski}
2554: J.~Polchinski, ``String Theory. Vol. 2: Superstring Theory And
2555: Beyond,'', Cambridge University Press, 1998.
2556:
2557: \bibitem{kutasov2}
2558: D.~Kutasov and V.~Niarchos, ``Tachyon Effective Actions In Open
2559: String Theory,'' arXiv:hep-th/0304045.
2560: %%CITATION = HEP-TH 0304045;%%
2561:
2562: \bibitem{larsen}
2563: F.~Larsen, A.~Naqvi and S.~Terashima, ``Rolling tachyons and
2564: decaying branes,'' arXiv:hep-th/0212248.
2565: %%CITATION = HEP-TH 0212248;%%
2566:
2567: \bibitem{hashimoto}
2568: K.~Hashimoto, P.~M.~Ho and J.~E.~Wang, ``S-brane actions,''
2569: arXiv:hep-th/0211090.
2570: %%CITATION = HEP-TH 0211090;%%
2571:
2572: \bibitem{ozetal}
2573: P.~Brax, G.~Mandal and Y.~Oz, ``Supergravity description of
2574: non-BPS branes,'' Phys.\ Rev.\ D {\bf 63}, 064008 (2001)
2575: [arXiv:hep-th/0005242].
2576: %%CITATION = HEP-TH 0005242;%%o
2577:
2578: \bibitem{nunez}
2579: J.~M.~Maldacena and C.~Nunez, ``Towards the large N limit of pure
2580: N = 1 super Yang Mills,'' Phys.\ Rev.\ Lett.\ {\bf 86}, 588
2581: (2001) [arXiv:hep-th/0008001].
2582: %%CITATION = HEP-TH 0008001;%%
2583:
2584: \bibitem{peet}
2585: A.~W.~Peet, ``TASI lectures on black holes in string theory,''
2586: arXiv:hep-th/0008241.
2587: %%CITATION = HEP-TH 0008241;%%
2588:
2589: \bibitem{andycft}
2590: A.~Strominger, ``Inflation and the dS/CFT correspondence,'' JHEP
2591: {\bf 0111}, 049 (2001) [arXiv:hep-th/0110087].
2592: %%CITATION = HEP-TH 0110087;%%
2593:
2594: \bibitem{us}
2595: F.~Leblond, D.~Marolf and R.~C.~Myers, ``Tall tales from de Sitter
2596: space. I: Renormalization group flows,'' JHEP {\bf 0206}, 052
2597: (2002) [arXiv:hep-th/0202094].
2598: %%CITATION = HEP-TH 0202094;%%
2599:
2600: \bibitem{martinec}
2601: J.~A.~Harvey, D.~Kutasov and E.~J.~Martinec, ``On the relevance of
2602: tachyons,'' arXiv:hep-th/0003101.
2603: %%CITATION = HEP-TH 0003101;%%
2604:
2605: \end{thebibliography}
2606: %====================================================================+
2607: \end{document}
2608: