hep-th0207105/near
1: \documentclass[12pt]{article}
2: \input epsf.sty
3: \topmargin -.5cm
4: \textheight 21cm
5: \oddsidemargin -.125cm
6: \textwidth 16cm
7: 
8: 
9: 
10: \newcommand{\omg}{\omega}
11: \newcommand{\ff}{F}
12: \newcommand{\eps}{\epsilon}
13: \newcommand{\ra}{\rangle}
14: \newcommand{\la}{\langle}
15: \newcommand{\T}{\chi_{T}(k)}
16: \newcommand{\Tm}{\chi_{T}(k')}
17: \newcommand{\Cn}{{\cal C}_n}
18: \newcommand{\vp}{\varphi}
19: \newcommand{\ve}{\varepsilon}
20: \newcommand{\tl}{\wt\lambda}
21: \newcommand{\hl}{\wh\lambda}
22: 
23: \newcommand{\vv} {\bar v}
24: \newcommand{\uu} {\bar u}
25: \newcommand{\K}{{\rm K_1}}
26: \newcommand{\Kt}{{\rm \widetilde K_1}}
27: 
28: 
29: \newcommand{\B}{b'}
30: \newcommand{\C}{c'}
31: \newcommand{\bB}{\bar b'}
32: \newcommand{\bC}{\bar c'}
33: \newcommand{\Bu}{B_{\vec u}}
34: \newcommand{\VV}{{\cal V}}
35: \newcommand{\BB}{{\cal B}}
36: \newcommand{\II}{{\cal I}}
37: \newcommand{\GG}{{\cal G}}
38: \newcommand{\FF}{{\cal F}}
39: \newcommand{\HH}{{\cal H}}
40: \newcommand{\MM}{{\cal M}}
41: \newcommand{\CC}{{\cal C}}
42: \newcommand{\OO}{{\cal O}}
43: \newcommand{\QQ}{{\cal Q}}
44: \newcommand{\PP}{{\cal P}}
45: \newcommand{\EE}{{\cal E}}
46: \newcommand{\LL}{{\cal L}}
47: \newcommand{\lll}{\langle\langle}
48: \newcommand{\rrr}{\rangle\rangle}
49: \newcommand{\square}{\Box}
50: \newcommand{\half}{{1\over 2}}
51: \newcommand{\wt}{\widetilde}
52: \newcommand{\wh}{\widehat}
53: \newcommand{\wc}{\check}
54: \newcommand{\wb}{\bar}
55: \newcommand{\bd}{\bar{\rm D}}
56: \newcommand{\RR}{{\cal R}}
57: \newcommand{\NN}{{\cal N}}
58: \newcommand{\TT}{{\cal T}}
59: 
60: \newcommand{\be}{\begin{equation}}
61: \newcommand{\ee}{\end{equation}}
62: \newcommand{\ben}{\begin{eqnarray}\displaystyle}
63: \newcommand{\een}{\end{eqnarray}}
64: \newcommand{\refb}[1]{(\ref{#1})}
65: \newcommand{\p}{\partial}
66: \newcommand{\sectiono}[1]{\section{#1}\setcounter{equation}{0}}
67: \renewcommand{\theequation}{\thesection.\arabic{equation}}
68: %\renewcommand{\theequation}{\arabic{equation}}
69: 
70: 
71: 
72: \begin{document}
73: {}~
74: \hfill\vbox{\hbox{hep-th/0207105}%\hbox{MRI-P-020703}
75: }\break
76: 
77: \vskip .6cm
78: 
79: 
80: \centerline{\Large \bf
81: Time Evolution in Open String Theory
82: }
83: 
84: 
85: \medskip
86: 
87: \vspace*{4.0ex}
88: 
89: \centerline{\large \rm 
90: Ashoke Sen }
91: 
92: \vspace*{4.0ex}
93: 
94: \centerline{\large \it Harish-Chandra Research
95: Institute}
96: 
97: \centerline{\large \it  Chhatnag Road, Jhusi,
98: Allahabad 211019, INDIA}
99: 
100: \centerline {and}
101: \centerline{\large \it Department of Physics, Penn State University}
102: 
103: \centerline{\large \it University Park,
104: PA 16802, USA}
105: 
106: \centerline{E-mail: asen@thwgs.cern.ch, sen@mri.ernet.in}
107: 
108: \vspace*{5.0ex}
109: 
110: \centerline{\bf Abstract} \bigskip 
111: 
112: We discuss a general iterative procedure for constructing time dependent
113: solutions in open string theory describing rolling of a generic tachyon
114: field away from its maximum. These solutions are characterized by two
115: parameters labelling the initial position and velocity of the tachyon
116: field, but one of these parameters can be removed by using time
117: translation symmetry.  The Wick rotated version of the resulting one
118: parameter family of inequivalent solutions describes a one parameter
119: family of boundary conformal field theories, each member of which is
120: related to the boundary conformal field theory describing the original
121: D-brane system by a nearly marginal deformation. We apply this technique
122: to construct a time dependent solution on a D-brane in bosonic string
123: theory which can be interpreted as the creation of a lower dimensional
124: brane during the decay of an unstable D-brane. 
125: 
126: \vfill \eject
127: 
128: \baselineskip=16pt
129: 
130: \tableofcontents
131: 
132: 
133: \sectiono{Introduction and Summary} \label{s1}
134: 
135: Time dependent solutions in string theory have received attention 
136: recently, both in closed string theory\cite{closed} and in open 
137: string 
138: theory\cite{0202210,0203211,0203265,0204203,0205085,0205098,0206102,0207004}.
139: (For early studies of open string tachyons see \cite{old}.)
140: In
141: particular in 
142: open string 
143: theory a 
144: special class of time 
145: dependent solutions were constructed which describe the rolling of the 
146: tachyon field on a non-BPS D-brane\cite{9806155,9809111}, or a 
147: brane-antibrane pair away from the 
148: maximum of the potential. Possible applications of these solutions to 
149: cosmology have been discussed in \cite{cosmo}.\footnote{For earlier 
150: attempts to 
151: apply the rolling tachyon solution to cosmology, see \cite{earl}.}
152: 
153: One common feature of the solutions associated with the rolling of an open 
154: string tachyon is that after removing the trivial 
155: parameter of the solution associated with time translation, the solutions 
156: are characterized by one parameter labelling the energy of the
157: solution.
158: This is precisely what one
159: would expect in a scalar field theory
160: with standard kinetic term, where a general spatially 
161: homogeneous solution 
162: will be characterized by the initial position and velocity of the scalar 
163: field, and one of these two parameters can be removed by using time 
164: translation invariance which allows us to set either the initial position 
165: or the initial velocity to zero.\footnote{In a multi-well potential there 
166: can be more than one
167: family of inequivalent solutions labelled by the energy.}
168: However this is somewhat surprising in 
169: open string field theory where the interaction terms have infinite number 
170: of derivatives.\footnote{Of course while the analysis of 
171: \cite{0203211,0203265} generates 
172: a one parameter family of time dependent solutions, it does not rule out 
173: the existence of other solutions and so it is still conceivable that the 
174: full theory has more solutions.}
175: 
176: The analysis of \cite{0203211,0203265} 
177: was facilitated by the fact that the rolling 
178: tachyon solution was related by Wick rotation to a boundary conformal 
179: field theory (BCFT) that was obtained from the original D-brane system 
180: by an 
181: {\it exactly marginal} deformation. The deformation parameter was related 
182: to 
183: the parameter labelling the inequivalent solutions. We cannot hope that 
184: this will be true for 
185: more general tachyonic states which might be present on a generic D-brane 
186: system, {\it e.g.} tachyons on an intersecting brane 
187: system\cite{9704006,9703217}.\footnote{Possible application of rolling
188: tachyon on 
189: intersecting D-brane system to cosmology has been emphasized in
190: \cite{inter}.} Thus 
191: the question arises: is it still possible to construct a one 
192: parameter family of inequivalent solutions which describe the rolling of a 
193: generic 
194: open string tachyon away from its maximum? 
195: 
196: This is the question we address 
197: in the paper, and show that under certain generic conditions, the answer 
198: to this question is in the affirmative. In particular once these 
199: conditions are satisfied, we have a one parameter family of inequivalent 
200: solutions, which are related by Wick rotation to a one parameter family of 
201: euclidean BCFT's. However, these BCFT's are not related to each other by a 
202: a marginal deformation. Instead, each of them is related to the BCFT 
203: describing the original unstable D-brane system by a nearly marginal 
204: deformation. The nearly marginal operator, however, is 
205: different for 
206: different solutions.
207: 
208: Unfortunately, although our analysis establishes the existence of a one 
209: parameter family of rolling tachyon solutions for a generic open string 
210: tachyon, unlike in the cases discussed in \cite{0203211,0203265},
211: generically the 
212: deformed BCFT's are 
213: not exactly solvable. As a result, we cannot give an explicit construction 
214: of 
215: the energy-momentum tensor by Wick rotating the boundary state of the 
216: deformed theory. Nevertheless, the success of this approach in generating 
217: explicit
218: solution in special cases where the deformation is exactly 
219: marginal\cite{0203211,0203265}, as well as in generating general 
220: class of solutions in 
221: $p$-adic string theory\cite{mo-zw} leads us to believe that this approach 
222: has a more 
223: general validity.
224: 
225: %The rest of the paper is organised as follows. 
226: 
227: In section \ref{s2} we 
228: illustrate our 
229: method of constructing the rolling tachyon solution by studying the 
230: example of a 
231: scalar field theory with standard kinetic term and potential $V(\phi)$ 
232: with a maximum at $\phi=0$. The idea is to relate the 
233: rolling tachyon solution via Wick rotation to the equation of motion for a 
234: scalar field with potential $-V(\phi)$. We can now solve for the periodic 
235: motion of this scalar field around $\phi=0$ using perturbation 
236: theory where we let the period of oscillation depend on the amplitude, and 
237: solve for the period and the orbit as a perturbation series in the 
238: amplitude\cite{golds}. Once we have obtained a solution this way we can 
239: inverse Wick 
240: rotate the solution to find a rolling tachyon solution. The amplitude of 
241: oscillation in the Wick rotated theory labels the initial
242: value of the 
243: rolling tachyon.
244: 
245: In section \ref{s3} we show how under certain conditions this procedure
246: can be generalized to open string field theory to construct rolling
247: tachyon solutions associated with a generic tachyon field. 
248: We also generalize this construction to the case where multiple tachyons
249: roll simultaneously beginning with arbitrary initial position and
250: velocity. We reinterpret
251: this construction in section \ref{s4} in terms of boundary conformal field
252: theory. In particular we show that in the Wick rotated theory, the family
253: of solutions labelled by the initial position of the tachyon field 
254: correspond 
255: to a family of BCFT's, and each member of this family
256: is
257: related to the BCFT describing the original D-brane by a nearly marginal
258: deformation. 
259: 
260: In section \ref{s5} we consider a specific example -- that of
261: a D-$p$-brane in bosonic string theory with one direction compactified on 
262: a circle of radius $R$,
263: -- and consider the rolling of the $(p-1,1)$ dimensional tachyon, obtained
264: by taking the first momentum mode of the $(p,1)$ dimensional tachyon along
265: the circle. We show how in the Wick rotated theory we can construct a one
266: parameter family of boundary conformal field theories, each related to the
267: original D-$p$-brane by a nearly marginal deformation, and how the inverse
268: Wick rotation of these BCFT's generate the family of rolling tachyon
269: solutions.  Unfortunately these BCFT's are not exactly solvable and hence
270: we cannot compute the analytic expressions for the energy-momentum tensor 
271: associated with these solutions.
272: However at a particular value of the radius, $R=\sqrt 2$, the family of 
273: BCFT's become related by a marginal deformation and are exactly solvable. 
274: As a result we can compute the time evolution of the energy momentum 
275: tensor explicitly by taking the inverse Wick rotation of the boundary 
276: states associated with these BCFT's. We find that during the course of 
277: time evolution there is non-trivial flow of energy density along the 
278: compact direction, and at a certain finite value of 
279: time, the energy density at a particular location on this circle
280: blows up. More specifically, the energy density as a function of the
281: coordinate $y$ along the compact circle approaches a delta function
282: singularity as we approach this critical time.
283: As we pass this  critical time, there is an apparent loss of energy
284: density at the location of the singularity, with the total amount of
285: energy lost being equal to the total initial energy of the system. Thus
286: just after the critical time the energy density averages to zero. If we
287: naively continue the
288: formula for energy density beyond this critical time, the energy density
289: gradually evolves to zero
290: everywhere. We suggest a natural interpretation of this as the creation of
291: a codimension one lump from the decay of the original brane.
292: 
293: In section \ref{s6} we generalize the construction to the case of 
294: superstring theory. In particular we consider rolling tachyon solution 
295: that should describe the creation of a D-$(p-1)$ -- $\bar{\rm D}$-$(p-1)$ 
296: pair due to the decay of a non-BPS D-$p$-brane wrapped on a circle. As in
297: the case of bosonic 
298: string theory, we can construct 
299: this solution in the Wick rotated theory as a nearly marginal deformation 
300: of the original D-$p$-brane conformal field theory. However, this BCFT is 
301: not 
302: exactly solvable, and hence we cannot explicitly construct the 
303: space-time dependent energy 
304: momentum tensor associated with this solution.
305: In section \ref{s7} we discuss 
306: possible generalization of this method to the study of closed string
307: tachyon condensation. Although 
308: part of the argument can be generalized to the case of closed strings, the 
309: main obstacle to finding the solution in closed string theory arises 
310: due to our inability to solve the equations of motion of the graviton and 
311: the dilaton field in the background of rolling tachyon.
312: We conclude in section \ref{s8} with a few remarks.
313: 
314: \sectiono{Example in Scalar Field Theory} \label{s2}
315: 
316: In this section we shall discuss the general method for constructing time 
317: dependent solutions in the context of a scalar field theory, which we 
318: shall 
319: generalize to string field theory in the later sections. We begin 
320: with the action of a scalar field $\phi$ in $p+1$ dimensions with standard 
321: kinetic term and potential $V(\phi)$:
322: \be \label{e1}
323: S = - \int d^{p+1} x [ \p^\mu\phi \p_\mu\phi + V(\phi)]\, .
324: \ee
325: We shall further assume that $V(\phi)$ has a maximum at $\phi=0$, with
326: \be \label{e2}
327: V''(0) = -m^2\, .
328: \ee
329: We want to study time dependent solution of this equation determined by a
330: given initial condition. For simplicity let us consider spatially
331: homogeneous field configurations. In this case the solution is
332: characterized by two parameters, -- the initial position and velocity of
333: $\phi$. For definiteness we shall further restrict to solution with
334: total energy $E< V(0)$. In this case $\p_0\phi$ vanishes at some instant
335: of time when $V(\phi)=E$. We shall take this to be the origin of $x^0$. 
336: Thus the solution is now characterized by only one parameter $\lambda$, -- 
337: the value of $\phi$ at $x^0=0$. We shall be interested in the solution 
338: where $\lambda$ is small but not infinitesimal. 
339: 
340: 
341: The equation of motion is:
342: \be \label{e3}
343: \p_0^2 \phi + V'(\phi) = 0\, .
344: \ee
345: It is of course straightforward to (numerically) integrate this equation, 
346: but 
347: this cannot be generilized to string field theory where the equations of 
348: motion contain infinite number of derivatives. 
349: We shall follow an indirect method that can be generalized in the context 
350: of 
351: string field theory. The basic idea is the same as the one followed in 
352: \cite{0203211,0203265}, namely we make a Wick rotation $x^0=i x$, and 
353: write eq.\refb{e3} as
354: \be \label{e4}
355: \p_x^2 \phi - V'(\phi) = 0\, .
356: \ee
357: If $\phi=f(x)$ is a solution of eq.\refb{e4}, then $\phi=f(-i x^0)$ will 
358: be a solution to \refb{e3}. Thus the aim is now to solve eq.\refb{e4} 
359: with the boundary condition $\phi=\lambda$, $\p_x\phi=0$ at $x=0$. 
360: This clearly can be thought of as a motion of a particle in potential 
361: $-V(\phi)$. Since $V(\phi)$ has a maximum at $\phi=0$, $-V(\phi)$ has a 
362: minimum at $\phi=0$, and for small $\lambda$, solution to the equation of 
363: motion 
364: will oscillate around zero. 
365: The period of oscillaton $T\equiv 2\pi/\omg$ is in general a function of 
366: $\lambda$.\footnote{An analytic expression of 
367: $T$ is
368: given by
369: $$ T = \sqrt 2 \, \int_{\phi_1}^{\phi_2} {d\phi
370:  \over \sqrt{V(\phi)-V(\lambda)}}\, ,
371: $$ 
372: where $\phi_1$ and $\phi_2$ are turning points where 
373: $V(\lambda)=V(\phi_1)=V(\phi_2)$. The initial value $\lambda$ of $\phi$ 
374: can be identified with either $\phi_1$ or $\phi_2$. However, we shall not 
375: use this formula.} Thus the 
376: solution can be expanded as
377: \be \label{e9}
378: \phi(x; \lambda) = \sum_{n=0}^\infty a_n \cos(n\omg x)\, .
379: \ee
380: 
381: We shall now solve \refb{e4} using perturbation 
382: theory\cite{golds}.
383: If we write
384: \be \label{e5a}
385: V(\phi) = -{1\over 2} m^2 \phi^2 +  V_{int}(\phi)\, ,
386: \ee
387: then \refb{e4} takes the form
388: \be \label{e5b}
389: (\p_x^2 + m^2) \phi -  V_{int}'(\phi) = 0\, .
390: \ee 
391: $ V_{int}(\phi)$ is of order $\phi^n$ with $n\ge 3$. Thus
392: for small $\lambda$ the solution behaves as
393: \be \label{e10}
394: \phi = 
395: \lambda\cos(mx) + \OO(\lambda^2)\, . 
396: \ee
397: Comparing this with \refb{e9} we get
398: \be \label{e11}
399: a_1 = \lambda + \OO(\lambda^2), \qquad
400: \omg = m + \OO(\lambda), \qquad 
401: a_n = \OO(\lambda^2) \quad \hbox{for} \quad n=0, n\ge 2\, .
402: \ee
403: {}From this we see that we can trade in the parameter $\lambda$ for 
404: the coefficient $a_1\equiv\hl$, and rewrite \refb{e11} as 
405: \be \label{e8}
406: a_1=\hl,  \qquad
407: \omg = m + \OO(\hl), \qquad
408: a_n = \OO(\hl^2) \quad \hbox{for} \quad n=0, n\ge 2\, .
409: \ee
410: We now substitute \refb{e9} into eq.\refb{e5b} to get
411: \be \label{e12}
412: \sum_{n=0}^\infty \Big({n^2\omg^2}-m^2\Big) \, a_n \cos(n\omg x)
413: = - V_{int}'\Big( \sum_{n=0}^\infty \, a_n \cos(n\omg x)\Big)\, .
414: \ee
415: The right hand side of this equation contains terms quadratic and higher 
416: order in $a_n$. Thus we can solve the equations iteratively as follows. 
417: The zeroeth order approximation is taken to be
418: \be \label{e13}
419: \omg=m, \qquad a_n=0 \quad \hbox{for} \quad n=0, \quad n\ge 2\, .
420: \ee
421: Also $a_1$ is set equal to $\hl$ to all orders.
422: We 
423: substitute the $k$-th order results for the $a_n$'s and $\omg$ on the 
424: right 
425: hand side of \refb{e12} to compute the coefficient of $\cos(n\omg x)$ 
426: for every $n$. 
427: Comparing this with the coefficient of $\cos(n\omg x)$ on the left hand 
428: side, 
429: we determine the $(k+1)$th order values of the $a_n$'s for $n=0$ and $n\ge 
430: 2$. On the other hand equating the coefficient of the $\cos(\omg x)$ 
431: terms on 
432: both sides, we determine the value of $\omg$ to $(k+1)$th 
433: order.\footnote{Actually this determines $\omg$ to order $\hl^k$, but 
434: since the coefficients $a_n$ are of order $\hl$ or higher, this determines 
435: the solution to order $\hl^{k+1}$.} This is 
436: possible since $a_1$ is set equal to $\hl$, and we are determining all the 
437: coefficients in terms of $\hl$.
438: 
439: This gives a solution to eq.\refb{e4}. Given this, we can now generate a 
440: solution to eq.\refb{e3} by making the substitution $x=-ix^0$. This gives
441: \be \label{e14}
442: \phi(x^0) = \sum_{n=0}^\infty a_n(\hl) \cosh(n\omg(\hl) x^0)\, ,
443: \ee
444: where the coefficients $a_n$ are the same as the ones determined by the 
445: previous method. This gives a family of solutions characterized by one 
446: parameter $\hl$. $\hl$ determines the initial value $\lambda$ of $\phi$, 
447: with the precise relation between $\hl$ and $\lambda$ being given by:
448: \be \label{e15}
449: \lambda = \phi(0) = \sum_{n=0}^\infty a_n(\hl)\, .
450: \ee
451: 
452: \begin{figure}[!ht]
453: \leavevmode
454: \begin{center}
455: \epsfysize=5cm
456: \epsfbox{fig1.eps}
457: \end{center}
458: \caption{The plot of a particular time dependent solution in a scalar 
459: field theory with potential $-\phi^2/2 + \phi^3/3$ for the choice
460: $\hl=.5$. This gives $\phi(x^0=0)\simeq 0.598122$. The top curve is the 
461: result of direct numerical integration with this boundary condition, while
462: the bottom curve is obtained 
463: by perturbative techniques discussed here with 60 harmonics. As we can see 
464: the two curves 
465: coincide for $-\tau/2 < x < \tau/2$ where $\tau$ is the period 
466: of oscillation, but the perturbation theory breaks down beyond this 
467: range.} \label{f1}
468: \end{figure}
469: 
470: 
471: For example, if we take the $\phi^3$ field theory with potential $-{1\over
472: 2}\phi^2 + {1\over 3}\phi^3$, then $V'_{int}(\phi) = \phi^2$. Using the
473: first order solution $\phi(x^0)=\hl\cos(\omega x^0)$ on the right hand
474: side of \refb{e12}, we get the solution to second order in $\hl$: 
475: \be \label{esecond}
476: a_1 = \hl, \qquad a_0 \simeq {1\over 2} \hl^2, \qquad a_2 \simeq -{1\over
477: 6}
478: \hl^2, \qquad \omega \simeq 1\, .
479: \ee
480: Note that $\omega$ remains constant at 1 since the Fourier expansion of
481: $(\hl\cos x)^2$ does not contain a term proportional to $\cos x$.
482: Substituting \refb{esecond} on the right hand side of \refb{e12} and
483: comparing coefficients of $\cos(\omega x)$ on both sides we get the
484: equation for $\omega$ at next order:
485: \be \label{eyyy1}
486: (\omega^2-1) \hl \simeq -{5\over 6} \hl^3 \qquad \to \qquad \omega^2
487: \simeq 1
488: -{5\over 6} \hl^2\, .
489: \ee
490: The coefficients $a_n$ to order $\hl^3$ are given by:
491: \be \label{eyyy2}
492: a_1 = \hl, \qquad a_0 \simeq {1\over 2} \hl^2, \qquad a_2 \simeq -{1\over
493: 6}
494: \hl^2, \qquad a_3 \simeq {1\over 48} \hl^3\, .
495: \ee
496: The explicit expression for the solution becomes complicated at higher
497: order, but we can evaluate it numerically. 
498: As illustrated in Fig.\ref{f1},
499: explicit numerical analysis in this $\phi^3$ field theory shows that this 
500: procedure generates the time dependent solution in this theory, describing 
501: rolling of $\phi$ from the maximum towards the local minimum of 
502: $V(\phi)$, quite accurately for half a
503: period of oscillation in either direction of $x^0$. 
504: Beyond this range the expansion \refb{e14} 
505: diverges and we must take recourse to analytic continuation in order to 
506: study the solution. 
507: 
508: \sectiono{Solution in String Field Theory} \label{s3}
509: 
510: We now consider some specific D-brane system in bosonic string theory with 
511: a tachyonic mode. Associated with the zero momentum tachyon state, there 
512: is a boundary operator $V_T$ of dimension $h<1$. In $\alpha'=1$ unit, the 
513: associated tachyon 
514: field has mass$^2$
515: \be \label{e2.1}
516: (h-1) \equiv -m^2\, .
517: \ee
518: If we displace the tachyon field $T$ a distance $\lambda$ away from its 
519: maximum, and let the system evolve in time, then to leading order in 
520: $\lambda$ the solution is given by 
521: $\lambda\cosh(mx^0)$.  In the Wick rotated theory this corresponds to 
522: $\lambda\cos(mx)$.
523: The corresponding solution of the linearizerd 
524: equations 
525: of motion in string field theory is given by
526: \be \label{e2.2}
527: \lambda\, \cos(mX(0)) \, V_T(0)c_1|0\ra\, .
528: \ee
529: This is a BRST invariant state of $L_0$ eigenvalue 0. 
530: 
531: 
532: The full string field theory equation of motion is given by:
533: \be \label{e2.2a}
534: Q_B |\Psi\ra + |\Psi*\Psi\ra = 0\, ,
535: \ee
536: where $|\Psi\ra$ is an open string state of ghost number 1, and
537: $Q_B$ is the BRST 
538: charge. We shall try to find a solution to eq.\refb{e2.2a} in the Siegel 
539: gauge
540: \be \label{e2.2b}
541: b_0|\Psi\ra = 0\, ,
542: \ee
543: where the equation takes the form:
544: \be \label{e2.2c}
545: L_0|\Psi\ra = -b_0 |\Psi*\Psi\ra\, .
546: \ee
547: Our goal will be to first construct a solution to \refb{e2.2c} that
548: extends \refb{e2.2} to higher orders in $\lambda$, and then carry
549: out
550: an inverse Wick rotation to  construct a time dependent solution to
551: eq.\refb{e2.2c} 
552: describing the rolling of the tachyon away from its maximum.
553: We begin by defining:
554: \be \label{e2.3}
555: |\phi_0\ra = \cos(\omg X(0)) \, V_T(0)\, c_1|0\ra\, ,
556: \ee 
557: where $\omg$ is a number that will be determined 
558: below.
559: We choose a set of linearly independent basis states $|\phi_r\ra$ ($0\le 
560: r < \infty$) in a
561: subspace $\HH_1$ of first quantized open string states, 
562: satisfying the following conditions:
563: \begin{enumerate}
564: \item Each $|\phi_r\ra$ has ghost number 1.
565: \item Each $|\phi_r\ra$
566: is symmetric under $X\to -X$, and carries $x$-momentum which 
567: is an
568: integer multiple of $\omg$.
569: \item All $|\phi_r\ra$ for $r\ge 1$ are orthogonal to $c_0|\phi_0\ra$.
570: \end{enumerate}
571: $\HH_1$ by definition is the subspace spanned by the basis states
572: $\{|\phi_r\ra, r\ge 0\}$ satisfying these properties, and
573: we shall look for a solution $|\Psi\ra$ of 
574: eq.\refb{e2.2c} in this subspace. We shall also require that when we 
575: express the solution $|\Psi\ra$ as a linear combination of the basis 
576: states $|\phi_r\ra$, the 
577: coefficient of $|\phi_0\ra$ in $|\Psi\ra$ is given by $\hl$; this in fact 
578: will be the definition of the parameter $\hl$. For later use, we shall 
579: define:
580: \be \label{edefpri}
581: |\chi\ra' = |\chi\ra - \NN^{-1} \la\phi_0|\chi\ra \, c_0|\phi_0\ra, \qquad 
582: \NN \equiv \la \phi_0|c_0|\phi_0\ra\, ,
583: \ee
584: for any state of $|\chi\ra$ ghost number 2. Physically $|\chi\ra'$ gives 
585: $|\chi\ra$ with its component along $c_0|\phi_0\ra$ removed.
586: 
587: The construction of the solution $|\Psi_{(n)}\ra$ to order $\hl^n$
588: proceeds as follows.
589: We define
590: \be \label{e2.4}
591: |\Psi_1\ra = \hl |\phi_0\ra\, ,
592: \ee
593: and  $|\Psi_p\ra$ iteratively by using the relation:
594: \be \label{e2.5}
595: |\Psi_p\ra = -{b_0\over L_0} (|\Psi_{p-1}*\Psi_{p-1}\ra)' + \hl 
596: |\phi_0\ra\, .
597: \ee
598: We repeat \refb{e2.5} $(n-1)$-times 
599: to compute $|\Psi_n\ra$. 
600: If we define
601: \be \label{edeffn}
602: f_n(\omg,\hl) = \NN^{-1} \, \la \phi_0| \Psi_{n-1} * \Psi_{n-1}\ra\, ,
603: \ee
604: then we have
605: \be \label{e2.6}
606: (|\Psi_{n-1}*\Psi_{n-1}\ra)' = |\Psi_{n-1}*\Psi_{n-1}\ra - f_n(\omg,\hl) 
607: c_0 |\phi_0\ra\, .
608: \ee
609: We now define $\omg_n$ as the solution of the equation:
610: \be \label{e2.9}
611: \hl (\omg_n^2 - m^2)+
612: f_n(\omg_n,\hl) = 0\, ,
613: \ee
614: and take the $n$-th order solution $|\Psi_{(n)}\ra$ to be
615: \be \label{e2.11a}
616: |\Psi_{(n)}\ra = |\Psi_n\ra_{\omg=\omg_n}\, .
617: \ee
618: Note that $\omg$ at $n$th order is determined {\it after we have 
619: computed $f_n(\omg,\hl)$ by repeating the iterative procedure $(n-1)$ 
620: times}. 
621: Since $f_n(\omg,\hl)$ is of order $\hl^2$, \refb{e2.9} gives 
622: $\omg_n=m+\OO(\hl)$.\footnote{Actually from the definition of $\phi_0$ 
623: and 
624: conservation of $X$-momentum, it follows 
625: that $|\phi_0*\phi_0\ra$ contains states with $X$-momentum $\pm 2\omg$ 
626: or $0$, but does not contain a state of momentum $\pm\omg$. Thus 
627: $|\phi_0*\phi_0\ra$ does not contain any component along $c_0|\phi_0\ra$. 
628: As a result the order $\hl^2$ contribution to $f_n(\omega,\hl)$ vanishes,
629: and 
630: $f_n(\omega,\hl)=\OO(\hl^3)$. Thus $\omg=m+\OO(\hl^2)$. Generalising this 
631: argument we can see that 
632: $f_n(\omega,\hl)$ must be an odd function of $\hl$, and hence $\omg$ must
633: be 
634: an even function of $\hl$.}
635: 
636: We shall now show that $|\Psi_{(n)}\ra$ determined this way 
637: satisfies:
638: \be \label{e2.11}
639: L_0 |\Psi_{(n)}\ra = - b_0 |\Psi_{(n)}*\Psi_{(n)}\ra + 
640: \OO(\hl^{n+1})\, ,
641: \ee
642: and hence satisfies the string field theory equations of motion to order
643: $\hl^n$.
644: To do this, we
645: first note that \refb{e2.5}, \refb{e2.6} for $p=n$ gives
646: \be \label{e2.7}
647: L_0|\Psi_n\ra = -b_0 |\Psi_{n-1}*\Psi_{n-1}\ra +   
648: f_n(\omg,\hl)|\phi_0\ra  + \hl (\omg^2 - m^2) |\phi_0\ra\, ,
649: \ee
650: where we have used the fact that
651: \be \label{e2.8}
652: L_0 |\phi_0\ra = (\omg^2 - m^2) |\phi_0\ra\, .
653: \ee
654: Eq.\refb{e2.9}, \refb{e2.7} now give 
655: \be \label{e2.10} 
656: [L_0 |\Psi_n\ra + b_0 |\Psi_{n-1}*\Psi_{n-1}\ra]_{\omg=\omg_n}=0\, .
657: \ee 
658: We shall now show that 
659: \be \label{e2.12}
660: |\Psi_p\ra = |\Psi_{p-1}\ra + \OO(\hl^p)\, ,
661: \ee 
662: for any $p\le n$.
663: In that case if we replace $|\Psi_{n-1}\ra$ on 
664: the right hand side of \refb{e2.10} by $|\Psi_n\ra$, the net error will be 
665: of order $\hl^{n+1}$, since $\Psi_{n}$ itself is order $\hl$. 
666: This, together with \refb{e2.11a}, would establish \refb{e2.11}.
667: 
668: In order to establish \refb{e2.12} we use proof by induction.
669: We use \refb{e2.5} to get:
670: \be \label{e2.13}
671: |\Psi_{p+1}\ra - |\Psi_{p}\ra 
672: = -{b_0\over L_0} (|\Psi_{p}*\Psi_{p}\ra - 
673: |\Psi_{p-1}*\Psi_{p-1}\ra)'\, .
674: \ee
675: Now suppose that \refb{e2.12} holds for some value of $p$. In that case 
676: the 
677: right hand side of \refb{e2.13} 
678: is of 
679: order 
680: $\hl^{p+1}$, since
681: $|\Psi_{p-1}\ra$ itself is of order $\hl$. Since \refb{e2.12} holds 
682: for $p=1$, this proves \refb{e2.12} for 
683: all 
684: $p\le n$.
685: 
686: Note however that for the above argument to go through, we need to ensure 
687: that $b_0(|\Psi_{p}*\Psi_{p}\ra -
688: |\Psi_{p-1}*\Psi_{p-1}\ra)'$ does not have a state with $L_0$ eigenvalue 
689: of order $\hl$, since this will give an additional power of $\hl$ in the 
690: denominator from the operation of $1/L_0$ and will destroy the counting of 
691: powers of $\hl$ that we have used. It is precisely due to this reason that 
692: we have removed the component of $|\Psi_{p}*\Psi_{p}\ra$ along 
693: $c_0|\phi_0\ra$ in this expression, since $c_0|\phi_0\ra$ has $L_0$ 
694: eigenvalue 
695: $(\omg^2 - m^2)\sim \hl$. For our analysis to hold it is important that
696: besides $|\phi_0\ra$,  
697: there are no other states appearing in the computation of 
698: $|\Psi_{p}*\Psi_{p}\ra$, which has $L_0$ eigenvalue of order $\hl$ or 
699: less.
700: 
701: This requirement can be reformulated purely in terms of matter sector 
702: vertex operators as follows.
703: The possible problematic states appearing in the expression for
704: $|\Psi_{p}*\Psi_{p}\ra$ are ghost number 2 states which are not
705: annihilated by $b_0$, and has $L_0$ eigenvalue 0 in the $\hl\to 0$ limit.  
706: Such states are of the form: 
707: \be \label{ef1} 
708: O(0)\, c_0 c_1 |0\ra\, , 
709: \ee
710: where $O$ is a matter sector vertex operator whose conformal weight
711: approaches 1 as $\hl\to 0$. As long as the theory does not contain any
712: operator of this type which can appear in the operator product of
713: $\cos(\omg X) V_T$ with itself (other than $\cos(\omg X) V_T$ itself,
714: whose effect has already been taken into account) the analysis given above
715: remains valid.  Matter operators of the form $\p X$ of dimension 1, which 
716: could cause potential problems, do not do so since the $X\to -X$ symmetry 
717: prevents their appearance in 
718: the operator product of $\cos(\omg X) V_T$ with itself.
719: 
720: The situation simplifies if not even $c_0|\phi_0\ra$ appears in the 
721: computation of $|\Psi_p*\Psi_p\ra$ for $\omega=m$.
722: In this case we see from \refb{edeffn} that
723: $f_n(m,\hl)$ vanishes identically for arbitrary $\hl$. Thus 
724: eq.\refb{e2.9} can be solved by setting $\omg=m$. 
725: The cases analysed in refs.\cite{0203211,0203265} are of this type. 
726: 
727: Once we have obtained the solution in the Wick rotated theory, we can 
728: perform an inverse Wick rotation $x=-ix^0$ to obtain a time dependent 
729: solution in the Lorentzian theory. This amounts to replacing $\cos(\omg 
730: X)$ by $\cosh(\omg X^0)$, and the oscillator $\alpha_m$ of $X$ by 
731: $-i\alpha^0_m$ of $-iX^0$. Performing these operations on $\Psi_{(n)}$ we 
732: can generate a time dependent solution of the equations of motion to order 
733: $\hl^n$. The constant $\hl$ labelling the solution parametrizes the 
734: initial 
735: condition on the tachyon field. Of course, we expect the  series expansion 
736: to be valid only for a limited range of $x^0$, and for $x^0$ beyond this 
737: range we need to obtain the result via analytic continuation.
738: 
739: Since the discussion so far has been somewhat abstract, we shall now give 
740: explicit form of the solution to order $\hl^3$. We have:
741: \be \label{eexp1}
742: |\Psi_1\ra = \hl |\phi_0\ra\, ,
743: \ee
744: \be \label{eexp2}
745: |\Psi_2\ra = \hl |\phi_0\ra - \hl^2 \, {b_0\over L_0} \, 
746: |\phi_0*\phi_0\ra\, .
747: \ee
748: No subtraction is needed to this order since $|\phi_0*\phi_0\ra$ does not 
749: have a component along $c_0|\phi_0\ra$ due to $X$-momentum conservation. 
750: To next order\footnote{We could ignore the $\OO(\hl^4)$ term in 
751: $|\Psi_3\ra$, coming from the $*$-product of ${b_0\over L_0} \,
752: |\phi_0*\phi_0\ra$ with itself, in computing the solution to order
753: $\hl^3$.} \ben \label{eexp3}
754: |\Psi_3\ra &=& \hl |\phi_0\ra - \hl^2 \, {b_0\over L_0} |\phi_0 * 
755: \phi_0\ra \nonumber \\ 
756: && + {b_0\over L_0} \bigg( \hl^3 |\phi_0\ra * {b_0\over L_0} \,
757: |\phi_0*\phi_0\ra
758: + \hl^3 {b_0\over L_0} \, |\phi_0*\phi_0\ra * |\phi_0\ra + f_3(\omg,\hl)
759: c_0 
760: |\phi_0\ra \bigg) \nonumber \\ \cr
761: && + \OO(\hl^4)
762: \een
763: where
764: \ben \label{eexp4}
765: f_3(\omg,\hl) &=& -\NN^{-1} \hl^3 \la \phi_0| \bigg( |\phi_0\ra * 
766: {b_0\over 
767: L_0} \, |\phi_0*\phi_0\ra + {b_0\over L_0} \, |\phi_0*\phi_0\ra * 
768: |\phi_0\ra \bigg)\nonumber \\
769: &=& -2\, \NN^{-1} \hl^3 \la \phi_0 * \phi_0 | {b_0\over L_0} |\phi_0 * 
770: \phi_0\ra
771: \, .
772: \een
773: This can be related to the four point amplitude $A^{(4)}$ in string field 
774: theory involving four 
775: external states $|\phi_0\ra$. Taking into account the sum over $s$, $t$ 
776: and $u$ channel diagrams, and a factor of 2 coming from each vertex since 
777: the three point coupling in string field theory is accompanied by a factor 
778: of $1/3$ instead of $1/6$, we have\footnote{Throughout this paper 
779: $A^{(4)}$ will 
780: denote the amplitude in the euclidean string (field) theory.}
781: \be \label{eexp5}
782: A^{(4)}(\omg) = 12 \, \la \phi_0 * \phi_0 | {b_0\over L_0} |\phi_0 * 
783: \phi_0\ra \, .
784: \ee
785: Thus \refb{eexp4} can be written as
786: \be \label{eexp6}
787: f_3(\omg,\hl) = - {1\over 6} \NN^{-1} \hl^3 A^{(4)}(\omg)\, .
788: \ee
789: $\omg_3$ is then the solution to the equation
790: \be \label{eexp7}
791: (\omg_3^2 - m^2) = {1\over 6} \NN^{-1} \hl^2 A^{(4)}(\omg_3)\, .
792: \ee
793: To leading order we can set $\omg_3=m$ on the right hand side of 
794: \refb{eexp7} and get
795: \be \label{eexp8}
796: \omg_3 = \sqrt{ m^2 + {1\over 6} \NN^{-1} \hl^2 A^{(4)}(m)}\, .
797: \ee
798: $|\Psi_3\ra$ given in \refb{eexp3}, evaluated at $\omg=\omg_3$, then 
799: gives the solution $|\Psi_{(3)}\ra$ to order $\hl^3$.
800: 
801: This concludes our discussion on generating the time dependent solution of 
802: string field theory, describing the rolling of a tachyon away from its 
803: maximum. We shall end this section by mentioning three generalizations of
804: this analysis:
805: \begin{itemize}
806: \item We have discussed the case where the tachyon begins rolling from an 
807: initial configuration where its time derivative vanishes and the field is 
808: displaced from its maximum. In this case the total energy of the system is 
809: less than that at the maximum. We could also consider the case where we 
810: begin with a configuration where the tachyon field is at its maximum and 
811: has a small velocity. The leading order solution in this case 
812: is proportional to $\sinh(\omg x^0)$ with $\omg=m$. As discussed in 
813: \cite{0203211,0203265}, 
814: the full solution in this case can be found by making the replacement:
815: \be \label{ereplace}
816: x^0 \to x^0 + i \pi/(2\omg), \qquad  \hl \to - i \hl\, ,
817: \ee
818: in the solution we have obtained earlier.
819: \item 
820: The method discussed here 
821: can also be used to generate time dependent solutions describing 
822: oscillation of positive 
823: mass$^2$ fields about the minimum of their potential. In this case there 
824: is no need for Wick rotation. For a field $\phi$ of mass $M$, we take the 
825: first 
826: order solution to be
827: \be \label{ep1}
828: \hl\, \cos(\omg X^0(0)) \, V_\phi(0)c_1|0\ra\, ,
829: \ee
830: where $V_\phi$ is the vertex operator for the zero momentum scalar. 
831: At leading order, $\omg=M$.
832: We can now follow the iterative procedure described in this section to 
833: generate the 
834: solution to arbitrary power of $\hl$, and determine $\omg$ as a function 
835: of $\hl$ to that order. The $\hl$ dependence of $\omg$ will represent 
836: the anharmonicity of the oscillator due to the interaction terms in the 
837: string field theory action.
838: 
839: \item One could consider generalizing the construction to the case where
840: many tachyon fields roll simultaneously. In general, if we have $n$ scalar
841: fields then for a two derivative action we have $2n$ initial conditions.
842: We shall now show that even in string field theory we can construct a $2n$
843: parameter family of solutions describing the rolling of $n$ tachyons.
844: To leading
845: order the solution in the Wick rotated theory is:
846: \be \label{egena1}
847: |\Psi\ra = \sum_{i=1}^n \hl^{(i)} |\phi_0^{(i)}\ra\, ,
848: \ee
849: where
850: \be \label{egena2}
851: |\phi_0^{(i)}\ra = \cos(\omega^{(i)} X(0)+\vp^{(i)}) V_T^{(i)}(0)
852: c_1|0\ra\, .
853: \ee
854: $V_T^{(i)}$ is the vertex operator of the $i$th tachyon with
855: mass$^2=-m_{(i)}^2$, $\{\hl^{(i)}\}$ and $\{ \vp^{(i)}\}$ 
856: are the $2n$ parameters labelling the solution, and to leading order
857: $\omega^{(i)}=m^{(i)}$. We
858: shall assume that the $m^{(i)}$'s are incomensurate.
859: We
860: now generate the higher order solutions by iterating
861: \refb{egena1} as in the case of a single tachyon field, keeping
862: the coefficient of $|\phi^{(i)}_0\ra$ held fixed at $\hl^{(i)}$ to all
863: orders. The analog
864: of $|\chi\ra'$ for a state $|\chi\ra$ of ghost number 2 is defined as the
865: projection of $|\chi\ra$ in the subspace orthogonal to the one spanned by
866: the states $|\phi^{(i)}_0\ra$. Thus comparison of the coefficients of
867: $c_0|\phi^{(i)}_0\ra$ terms in the equation of motion determines
868: $\omega^{(i)}$ as a function of the $\hl^{(j)}$'s, whereas the
869: comparison of the other terms in the equation of motion determine the
870: component of $|\Psi\ra$ in the subspace orthogonal to the one spanned by
871: the $c_0|\phi^{(i)}_0\ra$'s. At the end of the computation we need to make
872: the replacement $X\to -iX^0$, and $\vp^{(i)}\to -i \theta^{(i)}$ where
873: $\theta^{(i)}$ are taken to be real parameters.
874: 
875: For this procedure to work we need to ensure that the operator product of
876: the operators $\cos(\omega^{(i)} X + \vp^{(i)}) V_T^{(i)}$ in the matter
877: sector does not generate another operator of dimension $=\OO(\hl^{(i)})$
878: outside the set $\{\cos(\omega^{(i)} X + \vp^{(i)})V_T^{(i)}\}$. The
879: generic dangerous operators are $\sin(\omega^{(i)} X +
880: \vp^{(i)})V_T^{(i)}$ and $\p
881: X$. The fact that $\sin(\omega^{(i)} X + \vp^{(i)})V_T^{(i)}$ is not
882: generated can be seen as follows.
883: If in a correlation
884: function we have one insertion of $\sin(\omega^{(i)} X + \vp^{(i)})$ and
885: multiple insertions of $\cos(\omega^{(j)} X + \vp^{(j)})$ for different
886: $j$, then we can evaluate this by writing the sines and cosines as sum of
887: exponentials. 
888: As long as the $\omega^{(j)}$'s are incomensurate,\footnote{This
889: restriction on $\omega^{(j)}$ means that the points in the parameter space
890: where perturbation theory fails are dense in the full parameter space, but
891: the situation here is no different from that in perturbation theory in
892: ordinary Hamiltonian system.} a correlator involving
893: $\exp(\pm i (\omega^{(j)}X + \vp^{(j)}))$ will be non-zero only if the
894: $\pm\omega^{(j)} X$ in the exponent cancel pairwise. As a result
895: the accompanying $\pm\vp^{(j)}$ also cancel pairwise. It is now easy to
896: see that for each term in the correlator involving the
897: $e^{i(\omega^{(i)} x + \vp^{(i)})}$ in $\sin(\omega^{(i)} x + \vp^{(i)})$,
898: there is an equal and opposite contribution involving the 
899: $e^{-i(\omega^{(i)} x + \vp^{(i)})}$ term in $\sin(\omega^{(i)} x +
900: \vp^{(i)})$, obtained by reversing the signs of the exponents of all the
901: other terms in the correlator.\footnote{There are many ways to pair terms
902: so that they cancel; here we only mention one of them.}
903: These two terms cancel. Thus such a
904: correlator vanishes, implying in turn that the operator product of
905: $\cos(\omega^{(j)} X + \vp^{(j)})$ does not contain a term proportional to
906: $\sin(\omega^{(i)} X + \vp^{(i)})$.
907: 
908: To see the absence of $\p X$ in the operator product of $\cos(\omega^{(j)}
909: X + \vp^{(j)})$, we consider a correlation function with one insertion of
910: $\p X$ and multiple insertions of various factors of $\cos(\omega^{(j)} X
911: + \vp^{(j)})$ on the boundary of the world-sheet. Since $\p X$ can be
912: thought of as the restriction of a bulk holomorphic current to the
913: boundary, this contribution can be expressed as a sum of poles of $\p X$
914: at the location of the various insertions of $\cos(\omega^{(j)} X +
915: \vp^{(j)})$. The residue of the pole at the insertion of
916: $\cos(\omega^{(k)} X + \vp^{(k)})$ is given by  
917: $\omega^{(k)}\sin(\omega^{(k)} X +
918: \vp^{(k)})$.
919: Thus
920: \ben \label{epp1}
921: &&\bigg\langle \p X(z) \prod_{s=1}^N  \cos(\omega^{(j_s)}
922: X(t_s)  + \vp^{(j_s)})\bigg\rangle \nonumber \\
923: &\propto&\sum_{r=1}^N {\omega^{(j_r)}\over (z-t_r)} \bigg\langle
924: \sin(\omega^{(j_r)}
925: X(t_r) + \vp^{(j_r)})
926: \prod_{s\ne r}  \cos(\omega^{(j_s)}
927: X(t_s) + \vp^{(j_s)})\bigg\rangle \, ,
928: \nonumber \\
929: && \qquad 1\le j_s \le n   
930: \een
931: The result vanishes by previous
932: argument. This, in turn, shows that the correlator of $\p X$ with a set of
933: $\cos(\omega^{(j)} X + \vp^{(j)})$ vanishes and hence we do not generate
934: $\p X$ in the operator product of $\cos(\omega^{(j)} X + \vp^{(j)})$.
935: 
936: Of course there may be other special dimension 1 operators which may be
937: generated in the operator product of $\cos(\omega^{(j)} X + \vp^{(j)})
938: V_T^{(j)}$. In that case the perturbative procedure outlined here for
939: generating solutions of string field theory equations of motion breaks
940: down.
941: 
942: \end{itemize}
943: 
944: \sectiono{Boundary Conformal Field Theory Description} \label{s4}
945: 
946: In this section we shall describe the BCFT associated with the solutions
947: constructed in the previous section. Since in the euclidean theory the
948: final solution is obtained by iterating an initial solution of the form
949: \be \label{e3.1} 
950: \hl \cos(\omg X(0)) \, V_T(0)\, c_1|0\ra\, , 
951: \ee 
952: we
953: might expect that this corresponds to a BCFT that is obtained by
954: perturbing the original BCFT by a boundary term
955: \be \label{e4.1} 
956: \tl \, \int dt
957: \cos(\omg X(t))\, V_T(t)\, , 
958: \ee 
959: where $\tl=\hl+\OO(\hl^2)$ and $t$ is a parameter labelling the boundary
960: of the world-sheet. In order for \refb{e4.1} to describe
961: a BCFT the $\beta$-function of the theory must vanish. Since $\cos(\omg
962: X)\, V_T$ has conformal weight $(\omg^2 + h)=(\omg^2-m^2+1)$, the
963: $\beta$ function for the coupling $\tl$ has the form: 
964: \be \label{e4.2}
965: \beta_{\tl}=(\omg^2-m^2)\tl + g(\omg, \tl)\, , 
966: \ee 
967: where
968: $g(\omg,\tl)$ represents higher order (in $\tl$) contribution to the
969: $\beta$-function. Thus the vanishing of the $\beta$ function 
970: requires:\footnote{See \cite{zam} for a discussion on this. For
971: application of the $\beta$-function equation to the study of tachyon
972: condensation, see \cite{bsft,0010247}.} 
973: \be \label{e4.3} 
974: (\omg^2-m^2)\tl + g(\omg, \tl) =0\, . 
975: \ee 
976: This is equivalent to
977: the equation \refb{e2.9} in string field theory. Of course, in order to
978: show that the perturbed theory is conformal we must ensure that the
979: $\beta$-functions associated with all other operators also vanish. To this 
980: end
981: we note that if $O$ denotes a boundary operator of dimension $h_O$, and if
982: we add this operator with coefficient $\lambda_O$ in the perturbed theory,
983: then the $\beta$-function of the coupling $\lambda_O$ is given by 
984: \be \label{e4.4} 
985: \beta_O = (h_O -1) \lambda_O + g_O(\tl, \lambda_O, \ldots)\, 
986: ,
987: \ee 
988: where $\ldots$ denote the coupling constants associated with the other
989: operators added to the theory, and $g_O$ is the contribution to $\beta_O$
990: from higher order quantum corrections. Since $g_O$ is of order $\tl^2$ or
991: higher, as long as $(h_O-1)$ is of order 1, we can make \refb{e4.4} vanish
992: by choosing a $\lambda_O$ of order $\tl^2$. This procedure breaks down if 
993: $(h_O-1)$ is of order $\tl$ or higher for some operator $O$ other than 
994: $\cos(\omg X) V_T$ (which has already been discussed in eq.\refb{e4.3}). 
995: Thus we require that the operator product of $\cos(\omg X) V_T$ with 
996: itself does not generate an operator of dimension $1+\OO(\tl)$ other
997: than $\cos(\omg X) V_T$. 
998: This is 
999: identical to the condition derived in the previous section for 
1000: obtaining 
1001: a solution of the string field theory equations of motion in a 
1002: perturbation expansion in $\tl$.
1003: 
1004: We shall now describe explicit computation of $\omg$ to order $\tl^2$.
1005: For this we can set $\omg=m$ in the second term in \refb{e4.3}. This 
1006: gives
1007: \be \label{egive}
1008: \omg^2 = m^2 - \tl^{-1} g(m,\tl)\, .
1009: \ee
1010: Thus
1011: we need to compute $g(m,\tl)$.
1012: For this computation we can pretend 
1013: that the $X$-coordinate is
1014: compactified with period $2\pi/m$. In that case $\cos(mX)V_T$, being a
1015: dimension 1 operator, represents a massless scalar field $\vp$ in this
1016: auxiliary
1017: theory. Let $\VV(\vp)$ denote the tree level effective potential for this
1018: scalar field 
1019: obtained after integrating out all the other fields. 
1020: (For $\omg\ne m$, there is also a 
1021: `mass term' ${1\over 2} (\omg^2-m^2)\vp^2$ coming from the 
1022: dimension of 
1023: the operator $\cos(\omg X)V_T$.)
1024: Then
1025: by the usual correspondence between the
1026: equations of motion of
1027: string theory and the condition of conformal invariance, we can compute
1028: the $\beta$-function $g(m,\tl)$ as
1029: \be \label{e5.4}
1030: g(m,\tl) = \VV'(\tl)\, ,
1031: \ee
1032: for suitably normalized $\VV$.
1033: In order to compute $\VV(\vp)$ we simply compute the tree level amplitude
1034: involving the massless scalars $\vp$ in this auxiliary string theory.
1035: First of
1036: all it is clear that there is no $\vp^3$ term in $\VV(\vp)$, since the
1037: three
1038: point amplitude of the scalar field described by the vertex operator
1039: $\cos(mX) V_T$ vanishes due to $X$-momentum conservation. Thus the leading
1040: contribution to $\VV(\vp)$ comes at the quartic order. The coefficient of 
1041: this term can be
1042: identified as $-(\NN^{-1}/4!)$ times the on-shell four point amplitude 
1043: $A^{(4)}$ 
1044: of the
1045: auxiliary scalar
1046: field at zero momentum:
1047: \be \label{e5.4a}
1048: \VV(\vp) = -(\NN^{-1}/4!) \, A^{(4)} \, \vp^4 + \OO(\vp^6)\, ,
1049: \ee
1050: where $\NN$ is the BPZ norm of the state $\cos(mX(0))V_T(0)|0\ra$ in the 
1051: matter sector.\footnote{The simplest way to keep track of the 
1052: normalization factor 
1053: $\NN$ is to first do the computation by assuming $\cos(mX)V_T$ to be a 
1054: normalized operator, and then account for the normalization by 
1055: replacing $A^{(4)}$ by $\NN^{-2} A^{(4)}$ and $\tl$ by $\NN^{1/2}\tl$ in 
1056: \refb{egg2}.}
1057: Thus we get, from \refb{e5.4},
1058: \be \label{egg1}
1059: g(m,\tl) = -(\NN^{-1}/6) \, A^{(4)} \tl^3 + \OO(\tl^5)\, ,
1060: \ee 
1061: and hence from \refb{egive}
1062: \be \label{egg2}
1063: \omg^2 = m^2 + {\NN^{-1}\over 6} \, A^{(4)}\, \tl^2 + \OO(\tl^4)\, .
1064: \ee
1065: This of course agrees with eq.\refb{eexp8} derived from string field 
1066: theory.
1067: 
1068: At this stage we should mention one subtle point. 
1069: Interpreting 
1070: the perturbed theory \refb{e4.1} as a deformed BCFT in the sense of
1071: renormalization group flow makes sense strictly 
1072: if the operator $\cos(\omg X) V_T$ is a relevant perturbation, {\i.e.} if 
1073: $\omg^2< m^2$. {}From \refb{e4.3}, \refb{e5.4} we see that this requires 
1074: $g(\omg, 
1075: \tl)$, or equivalently $\VV(\vp)$, to be positive near $\vp=0$. In the 
1076: analysis of 
1077: string field theory in section 
1078: \refb{s3} we did not encounter such a constraint. If $\VV$ turns out to be 
1079: negative, we can still find a solution in string field theory by choosing 
1080: $\omg^2-m^2$ to be positive. This will correspond to a net potential for 
1081: the auxiliary scalar field
1082: $\vp$ which has a minimum at $\vp=0$ and a maximum at a nearby point, and 
1083: the non-trivial solution of \refb{e4.3} will correspond to the maximum of 
1084: the potential. Thus even for negative $\VV$, we should expect a family of
1085: BCFT labelled by $\tl$, although members of this family are not obtained
1086: as deformation of the original BCFT by a relevant perturbation. Rather,
1087: the original BCFT is obtained as a relevant deformation of each member of
1088: this family.
1089: 
1090: When $\omega^2< m^2$ so that the perturbing operator is relevant, it is
1091: more natural to treat $\omega$ rather than $\lambda$ as the independent
1092: parameter. We simply deform the theory by the operator $\cos(\omega X)
1093: V_T$, and consider the boundary CFT to which the theory flows in the
1094: infrared. This will give a family of boundary CFT's labeled by $\omega$.
1095: Inverse Wick rotation of these boundary CFT's then give us the family of
1096: time dependent solutions corresponding to different initial conditions.
1097: 
1098: As special examples of the general class of perturbed BCFT's discussed 
1099: above, we can consider the case where
1100: $g(m, \tl)$ vanishes for arbitrary $\tl$, and as a 
1101: result \refb{e4.3} is satisfied for $\omg=m$ for all $\tl$. In this case 
1102: the 
1103: operator $\cos(mX)\, V_T$ is an exactly marginal operator. The cases 
1104: analysed in refs.\cite{0203211,0203265} are of this type.
1105: 
1106: Given a BCFT obtained by perturbing the euclidean theory by the operator
1107: \refb{e4.1}, we can construct a BCFT by the inverse Wick rotation of this
1108: perturbed theory. This will correspond to adding a perturbation
1109: \be \label{e4.5}
1110: \tl \cosh(\omg X^0(0)) \, V_T(0)\, c_1|0\ra\, ,
1111: \ee
1112: to the world-sheet theory with Minkowski signature space-time. The 
1113: boundary state\cite{boundary} associated with this BCFT, containing
1114: information about the 
1115: time evolution of the energy-momentum tensor\cite{em-tensor,0203265}, can
1116: be
1117: obtained from Wick 
1118: rotation of the boundary state associated with the euclidean theory. Time
1119: evolution of the sources of other massless closed string fields, {\it
1120: e.g.} the dilaton, the anti-symmetric tensor field $B_{\mu\nu}$, and
1121: various Ramond-Ramond fields, can also be extracted from the boundary
1122: state.
1123: 
1124: To see more explicitly how we can extract the energy-momentum tensor from
1125: the boundary state, we recall that\cite{0203265} if part of the
1126: boundary state 
1127: involving oscillators of level (1,1) acting on the ghost number 3 vacuum 
1128: has the form
1129: \be \label{e5.17}
1130: \int d^{26} k \, [\wt A_{\mu\nu}(k) \alpha^\mu_{-1}
1131: \bar\alpha^\nu_{-1} 
1132: + \wt B(k) (\bar
1133: b_{-1}
1134: c_{-1} + b_{-1} \bar c_{-1})] (c_0+\bar c_0)c_1\bar c_1 |k\ra\, ,
1135: \ee
1136: then the energy momentum tensor $T_{\mu\nu}$ is given by:
1137: \be \label{e5.18}
1138: T_{\mu\nu}(x) = K (A_{\mu\nu}(x) + \eta_{\mu\nu} B(x) )\, ,
1139: \ee
1140: where $A_{\mu\nu}(x)$ and $B(x)$ are Fourier transforms of $\wt
1141: A_{\mu\nu}(k)$ and $\wt B(k)$ respectively, and $K$ is a $\tl$-independent
1142: constant. 
1143: In the Wick rotated theory the energy momentum tensor computed using this
1144: procedure will have the form:
1145: \be \label{eem1}
1146: T_{\mu\nu}(x, \vec x) = \sum_{n=0}^\infty T^{(n)}_{\mu\nu}(\vec x)
1147: \cos(n\omega x)\, ,
1148: \ee
1149: where $\vec x$ denotes the spatial coordinates of the original theory.
1150: Then after Wick rotation we have:
1151: \ben \label{eem2}
1152: T_{00}(x^0, \vec x) &=& - \sum_{n=0}^\infty T^{(n)}_{xx}(\vec
1153: x)\cosh(n\omega
1154: x^0)\, , \nonumber \\
1155: T_{0i}(x^0, \vec x) &=& i \sum_{n=0}^\infty T^{(n)}_{xi}(\vec 
1156: x)\cosh(n\omega x^0)\, , \nonumber \\
1157: T_{ij}  &=& \sum_{n=0}^\infty T^{(n)}_{ij}(\vec
1158: x)\cosh(n\omega x^0)\, .
1159: \een
1160: Since in the Wick rotated theory the sum over $n$ in \refb{eem1} is
1161: expected to converge
1162: and generate a finite energy-momentum tensor, we expect that after Wick
1163: rotation the sum over $n$ in \refb{eem2} will converge for sufficiently
1164: small $x^0$ and $\tl$. We
1165: do expect the sum to diverge for large $x^0$, and checking
1166: whether the results for small $x^0$ can be analytically continued beyond
1167: the range of convergence or not, we shall have to determine whether during 
1168: the time
1169: evolution the energy momentum tensor hits a real singularity at a
1170: finite value of $x^0$. 
1171: 
1172: In principle, we should also be able to extract the energy-momentum tensor
1173: from the solution of the string field theory equations of motion
1174: constructed in the previous section, since this solution is expected to
1175: have all the information about the deformed boundary CFT. At present
1176: however it is not known how to do this.
1177: 
1178: \sectiono{Explicit Example} \label{s5}
1179: 
1180: In this section we shall discuss an explicit example of the construction 
1181: described in the previous section. The system that we shall consider is a 
1182: D-$p$-brane of bosonic string theory, with one direction wrapped on a 
1183: circle of radius $R>1$. 
1184: This has the usual tachyonic mode 
1185: of mass$^2=-1$, whose time evolution was discussed in \cite{0203211}. But 
1186: this also has another tachyonic mode of mass$^2$
1187: \be \label{em.1}
1188: R^{-2}-1\equiv -m^2
1189: \ee
1190: coming from 
1191: the 
1192: first momentum mode of the standard tachyon along the circle. If we denote 
1193: by $y$ the coordinate along the circle, then the matter sector vertex
1194: operator 
1195: associated with this tachyonic mode (with zero momentum along the
1196: non-compact directions) can be taken to be:
1197: \be \label{e5.1}
1198: V_T = \cos(Y/R)\, .
1199: \ee
1200: We shall discuss the rolling of this tachyon 
1201: when it is displaced from its maximum. As usual we denote by $x$ the Wick 
1202: rotated time coordinate. The conformal field theory 
1203: associated with the 
1204: (Wick rotated) rolling tachyon solution is obtained by perturbing the 
1205: original free conformal field theory by the boundary operator
1206: \be \label{e5.2}
1207: \tl\,  \int dt \, \cos(\omg X(t)) \, \cos(Y(t)/R)\, .
1208: \ee
1209: Eq.\refb{e4.3} now takes the form:
1210: \be \label{e5.3}
1211: (\omg^2 + R^{-2} - 1)\tl + g(\omg,\tl) = 0\, .
1212: \ee
1213: To leading order the solution to this equation is 
1214: $\omg=m=\sqrt{1-R^{-2}}$. 
1215: {}From eq.\refb{egg2} we see that the computation of $\OO(\tl^2)$ 
1216: correction to $\omega$ 
1217: requires us to compute $A^{(4)}$ involving four external states associated 
1218: with the matter sector vertex operator $\cos(m X) \cos(Y/R)$.
1219: This is the task to which 
1220: we shall now turn.
1221: 
1222: Expressing the scalar field vertex operator as
1223: \be \label{e5.5}
1224: \cos(mX) \cos(Y/R) = {1\over 4} (e^{imX}+e^{-imX}) \, 
1225: (e^{iY/R} + e^{-iY/R})\, ,
1226: \ee
1227: we can relate the four point amplitude $A^{(4)}$ of the scalar field to 
1228: the 
1229: Veneziano amplitude involving external states associated with 
1230: matter sector vertex operators $e^{\pm i m X} e^{\pm i Y/R}$.
1231: Since the right hand side of eq.\refb{e5.5} has four terms, there are
1232: altogether $4^4=256$ terms involved in the computation of the four point
1233: amplitude $A^{(4)}$. However all but 36 of them
1234: vanish by $X$- or $Y$-momentum conservation. 12 of the non-vanishing terms
1235: are each given by ${1\over 4^4} V(s=-4, t=0)$, and each of the rest 24
1236: terms is given by ${1\over 4^4} V(s=-4R^{-2}, t = -4(1 - R^{-2}))$, where
1237: $V(s,t)$ is the Veneziano amplitude
1238: \ben \label{e5.6}
1239: V(s,t) &=& 2\, \bigg( {\Gamma(-1-s) \Gamma(-1-t)\over \Gamma(-2-s-t)}
1240: + {\Gamma(-1-s) \Gamma(-1-u)\over \Gamma(-2-s-u)}
1241: +{\Gamma(-1-u) \Gamma(-1-t)\over \Gamma(-2-u-t)} \bigg), \nonumber \\ \cr
1242: && \qquad \qquad u=-4-s-t\, .
1243: \een
1244: Thus we get
1245: \ben \label{e5.7}
1246: A^{(4)} &=& {1\over 4^4} \, ( 12 V(s=-4, t=0) + 24 V(s=-4 R^{-2}, 
1247: t=-4(1-R^{-2})) \nonumber \\
1248: &=& -{3\over 32} \, (R^2 - 2) [ 4 R^{-2} \ff'(4R^{-2}) / \ff(4R^{-2}) + 
1249: \ff(4 
1250: R^{-2}) - 1]\, ,
1251: \een
1252: where
1253: \be \label{e5.8}
1254: \ff(x) = \Gamma(1-x) \, \Gamma(1+x) = {\pi x\over \sin(\pi x)}\, .
1255: \ee
1256: Using eqs.\refb{egg2}, \refb{e5.7}, and the fact that $\NN=1/4$ for the
1257: operator \refb{e5.5}, we now get
1258: \be \label{e5.10}
1259: \omg^2 = 1 - R^{-2} - {1\over 16} (R^2-2)  [ 4 R^{-2} \ff'(4R^{-2}) / 
1260: \ff(4R^{-2}) + \ff(4
1261: R^{-2}) - 1]\tl^2 + \OO(\tl^4)\, .
1262: \ee
1263: This determines the value of $\omg$ to order $\tl^2$ for which 
1264: perturbation by the vertex 
1265: operator $\cos(\omg X) \cos(Y/R)$ generates a boundary CFT. The time 
1266: dependent solution describing the rolling tachyon is then 
1267: determined by the inverse Wick rotation $X=-iX^0$ of this solution. In 
1268: particular by inverse Wick rotating the boundary state associated with the 
1269: perturbed euclidean BCFT, we can determine the energy-momentum tensor 
1270: associated with the time dependent solution.
1271: 
1272: Unfortunately for generic $R$ the perturbed BCFT is not solvable, and 
1273: hence 
1274: we cannot explicitly compute the boundary state associated with this BCFT. 
1275: For $R\to\infty$ the amplitude given in eq.\refb{e5.7} vanishes. This is 
1276: consistent with the fact that in this limit the perturbing operator 
1277: $\cos(X)$ is exactly marginal and hence the corresponding $\beta$-function 
1278: must vanish. The associated conformal field theory has been
1279: discussed in detail in \cite{marginal} and its inverse Wick rotation has
1280: been used
1281: in \cite{0203211,0203265} to describe the rolling of a spatially
1282: homogeneous tachyon field on the D-brane. 
1283: $A^{(4)}$ also vanishes at $R=1$ where the perturbing
1284: operator is $\cos(Y)$. There is another value of $R$
1285: where $A^{(4)}$ vanishes, 
1286: namely at $R=\sqrt 2$.\footnote{Note that $\ff(4R^{-2})$ diverges at 
1287: $R=\sqrt 2$ and so in evaluating \refb{e5.10} for $R=\sqrt{2}$ we need to 
1288: carefully take the limit. Nevertheless the right hand side of \refb{e5.10} 
1289: can be shown to vanish at $R=\sqrt 2$.}
1290: Thus one might expect that the corresponding perturbation becomes exactly 
1291: marginal at this point. To see that this is indeed the case, let us note 
1292: that at $\omega=1/\sqrt 2$, the perturbing operator is
1293: given by:
1294: \be \label{e5.11}
1295: \tl \, \int dt \, \cos(X(t)/\sqrt 2) \, \cos(Y(t)/\sqrt 2)
1296: = {\tl\over 2} \, \Big[\cos\Big((X(t) + Y(t))/\sqrt 2\Big) + 
1297: \cos\Big((X(t) 
1298: -
1299: Y(t))/\sqrt 2\Big)\Big]\, .
1300: \ee
1301: This corresponds to perturbing the theory by a sum of two commuting 
1302: exactly marginal 
1303: deformations, and hence represents a marginal deformation. The associated 
1304: BCFT is exactly solvable. In fact, defining
1305: \be \label{e5.12}
1306: Z^1 = (X+Y)/\sqrt 2 \, , \qquad Z^2 = (X-Y)/\sqrt 2\, ,
1307: \ee
1308: we see that the net boundary state is given by the product
1309: \be \label{e5.13}
1310: |\BB\ra_{Z^1} \otimes |\BB\ra_{Z^2} \otimes |\BB\ra_{c=24} \otimes 
1311: |\BB\ra_{ghost}\, .
1312: \ee
1313: $|\BB\ra_{c=24}$ denotes the boundary state associated with the 24 free 
1314: scalar fields, and $|\BB\ra_{ghost}$ denotes the boundary state associated 
1315: with the ghost fields. These are given by their free field expressions. On 
1316: the 
1317: other hand the relevant part of the boundary states $|\BB\ra_{Z^1}$ and $ 
1318: |\BB\ra_{Z^2}$, associated with the perturbed BCFT associated with the 
1319: scalar fields $Z^1$ and $Z^2$, can be read out from the results of 
1320: \cite{marginal,9811237,0108238,0203265}. If we consider for simplicity the
1321: case of 
1322: the 
1323: D-25-brane, the part of the boundary state, relevant for the computation 
1324: of the energy-momentum tensor, is proportional to:
1325: \ben \label{e5.14}
1326: && \bigg[f\Big((iX(0)+iY(0))/\sqrt 
1327: 2\Big) 
1328: - {1\over 2} (\alpha^X_{-1} + \alpha^Y_{-1}) (\bar\alpha^X_{-1} + 
1329: \bar\alpha^Y_{-1}) g\Big((iX(0)+iY(0))/\sqrt 2\Big) \bigg] \nonumber \\
1330: && \bigg[f\Big((iX(0)-iY(0))/\sqrt 2\Big)
1331: - {1\over 2} (\alpha^X_{-1} - \alpha^Y_{-1}) (\bar\alpha^X_{-1} -
1332: \bar\alpha^Y_{-1}) g\Big((iX(0)-iY(0))/\sqrt 2\Big) \bigg] \nonumber \\
1333: && \bigg[1-\alpha^i_{-1}\bar\alpha^i_{-1}\bigg]  \, \bigg[1-\bar 
1334: b_{-1}
1335: c_{-1} - b_{-1} \bar c_{-1}\bigg] (c_0+\bar c_0)c_1\bar c_1 |0\ra\, .
1336: \een
1337: Here $\alpha^i_{-1}$, $\bar\alpha^i_{-1}$ for $2\le 
1338: i\le 25$ are the 
1339: oscillators associated with the spectator bosons $X^i$,
1340: $\alpha^X_{-1}$, $\bar\alpha^X_{-1}$, $\alpha^Y_{-1}$, $\bar\alpha^Y_{-1}$ 
1341: are the oscillators associated with the scalars $X$ and $Y$, and $b_n$,
1342: $c_n$, $\bar b_n$, $\bar c_n$ are the oscillators associated with the
1343: ghost fields $b$, $c$, $\bar b$, $\bar c$.
1344: The functions $g(x)$ and $f(x)$ have been given in \cite{0203211,0203265}:
1345: \be \label{e5.15}
1346: f(x^0)={1\over 1 + e^{x^0} \sin(\tl\pi/2)} + {1 \over
1347: 1 + e^{-x^0} \sin(\tl\pi/2)} - 1\, , \qquad
1348: g(x^0) = \cos(\tl\pi) +1 - f(x^0)\, .
1349: \ee
1350: Note that these formul\ae\ differ from that of \cite{0203211,0203265} by a 
1351: replacement $\tl\to \tl/2$, whose origin can be traced to the explicit 
1352: factor of $1/2$ multiplying $\tl$ on the right hand side of \refb{e5.11}.
1353: 
1354: We can now make an inverse Wick rotation $X=-iX^0$ to get the boundary 
1355: state in the Minkowski signature theory:
1356: \ben \label{e5.16}
1357: && \bigg[f\Big((X^0(0)+iY(0))/\sqrt 
1358: 2\Big) 
1359: - {1\over 2} (-i\alpha^0_{-1} + \alpha^Y_{-1}) (-i\bar\alpha^0_{-1} + 
1360: \bar\alpha^Y_{-1}) g\Big((X^0(0)+iY(0))/\sqrt 2\Big) \bigg] \nonumber \\
1361: && \bigg[f\Big((X^0(0)-iY(0))/\sqrt 2\Big)
1362: - {1\over 2} (-i\alpha^0_{-1} - \alpha^Y_{-1}) (-i\bar\alpha^0_{-1} -
1363: \bar\alpha^Y_{-1}) g\Big((X^0(0)-iY(0))/\sqrt 2\Big) \bigg] \nonumber \\
1364: && \bigg[1-\alpha^i_{-1}\bar\alpha^i_{-1}\bigg]  \, \bigg[1-\bar 
1365: b_{-1}
1366: c_{-1} - b_{-1} \bar c_{-1}\bigg] (c_0+\bar c_0)c_1\bar c_1 |0\ra\, .
1367: \een
1368: Comparing \refb{e5.16} with \refb{e5.17} we get,
1369: \ben \label{e5.19}
1370: A_{00} &=& {1\over 2} \, [g((x^0 + iy)/\sqrt 2) f((x^0 - i
1371: y)/\sqrt
1372: 2)  + g((x^0 - i 
1373: y)/\sqrt 
1374: 2) f ((x^0 + iy)/\sqrt 2)]\, , \nonumber 
1375: \\
1376: A_{0y} &=& A_{y0} = {i\over 2}  [g((x^0 + iy)/\sqrt 2) f((x^0 - i 
1377: y)/\sqrt 
1378: 2)  - 
1379: g((x^0 - i 
1380: y)/\sqrt
1381: 2)  f ((x^0 + iy)/\sqrt 2)]\, , \nonumber
1382: \\
1383: A_{yy} &=& -  {1\over 2} \,  [g((x^0 + iy)/\sqrt 2)f((x^0 - i
1384: y)/\sqrt
1385: 2) + g((x^0 - 
1386: i 
1387: y)/\sqrt
1388: 2)  f ((x^0 + iy)/\sqrt 2) ]\, , \nonumber
1389: \\
1390: A_{ij} &=& - \, \delta_{ij} \, 
1391: f ((x^0 + iy)/\sqrt 2)  f((x^0 - i
1392: y)/\sqrt
1393: 2)
1394: \, , \nonumber
1395: \\
1396: B &=& -  f ((x^0 + iy)/\sqrt 2)  f((x^0 - i
1397: y)/\sqrt
1398: 2)\, .
1399: \een
1400: Using \refb{e5.18}, \refb{e5.15} and \refb{e5.19} we get
1401: \ben \label{e5.20}
1402: T_{00} &=& K (A_{00} - B) = {K\over 2} (\cos(\tl\pi)+1) [f ((x^0 + 
1403: iy)/\sqrt 
1404: 2) + f((x^0 - i
1405: y)/\sqrt
1406: 2)]\, , \nonumber \\
1407: T_{0y} &=& K A_{0y} = {i K\over 2}  (\cos(\tl\pi)+1) [f((x^0 - i
1408: y)/\sqrt
1409: 2) - f ((x^0 +
1410: iy)/\sqrt
1411: 2) ] \nonumber \\
1412: T_{yy} &=&  K (A_{yy} + B) = -{K\over 2} (\cos(\tl\pi)+1) [f ((x^0 +
1413: iy)/\sqrt
1414: 2) + f((x^0 - i
1415: y)/\sqrt
1416: 2)]\, , \nonumber \\  
1417: T_{ij} &=& K(A_{ij} + B \delta_{ij})
1418: =  - 2 K\, \delta_{ij} \, f ((x^0 +
1419: iy)/\sqrt
1420: 2) \, f((x^0 - i
1421: y)/\sqrt
1422: 2)
1423: \een
1424: All other components of $T_{\mu\nu}$ vanish.
1425: {}From these expressions one can easily verify the conservation laws:
1426: \be \label{e5.21}
1427: \p_0 T_{00} - \p_y T_{0y} = 0, \qquad \p_0 T_{0y} - \p_y T_{yy} = 0\, .
1428: \ee
1429: If we consider a general D-$p$-brane with $p<25$, then $T_{ij}$ associated
1430: with the $(25-p)$ transverse directions vanish. Furthermore the
1431: non-vanishing components of $T_{\mu\nu}$ are each multiplied by a delta
1432: function involving the transverse coordinates. The $\tl$ independent
1433: constant $K$ is computed by requiring that for $\tl=0$, $T_{00}$ should
1434: reproduce the tension $\TT_p$ of a D-$p$-brane. This gives
1435: \be \label{ektp}
1436: K = {\TT_p\over 2}\, .
1437: \ee
1438: 
1439: {}From \refb{e5.15} we get
1440: \be \label{e5.22}
1441: f((x^0\pm i y)/\sqrt 2) = 
1442: {1\over 1 + e^{x^0/\sqrt 2} e^{\pm iy/\sqrt 2} \sin(\tl\pi/2)} + {1 \over
1443: 1 + e^{-x^0/\sqrt 2}  e^{\mp iy/\sqrt 2} \sin(\tl\pi/2)} - 1\, .
1444: \ee
1445: Thus $f((x^0\pm iy)/\sqrt 2)$ and hence
1446: the energy momentum tensor hits a singularity at
1447: \be \label{e5.23}
1448: x^0 = \sqrt 2 \, \ln \bigg|{1\over \sin(\tl\pi/2)}\bigg| \equiv x^0_c,
1449: \qquad y = \cases {\sqrt 2 \pi\, , \qquad \hbox{for} \quad \tl>0\, , \cr
1450: 0\, , \qquad \, \, \quad \, \hbox{for} \quad \tl<0\, .}
1451: \ee
1452: In particular the energy-density at the singular point approaches 
1453: $\infty$. Thus we cannot naively evolve the system beyond this time using
1454: the 
1455: classical open string field equations.
1456: 
1457: \begin{figure}[!ht]
1458: \leavevmode
1459: \begin{center}
1460: \epsfysize=8cm
1461: \epsfbox{fig2.eps}
1462: \end{center}
1463: \caption{The contour plot of the time evolution of the energy density
1464: $T_{00}$ computed from eq.\refb{e5.20}, \refb{e5.22} for $\tl=-0.1$. The
1465: horizontal axis
1466: is the $y$ axis, the vertical axis is the time ($x^0$) axis, and the
1467: shading at a
1468: given point $(y,t)$ shows the value of $T_{00}$, with darker shade
1469: representing higher energy density. We see from this that the $T_{00}$
1470: hits a singularity at $x^0\simeq 2.62$, $y=0$. At this space-time point, 
1471: the system
1472: loses a net amount of energy equal to the total initial energy of the
1473: system.  Immediately after this
1474: singularity the energy density becomes negative around $y=0$, as shown by 
1475: the lighter regions, and eventually the energy density approaches zero
1476: everywhere.} \label{f2} \end{figure}
1477: It is instructive to examine the nature of the singularity in a little 
1478: more detail. Using
1479: eqs.\refb{e5.20}, \refb{e5.22} we get the total energy of the system to 
1480: be:
1481: \be \label{etote}
1482: \EE = \int_0^{2\pi\sqrt 2} dy \, T_{00} = 2\pi\sqrt 2 K 
1483: (1 + \cos(\tl\pi))\, ,
1484: \ee
1485: for $x^0< x^0_c$. If we
1486: naively 
1487: continue the expression for $T_{00}$ to $x^0>x^0_c$, we find 
1488: that the total energy vanishes. In order to see the origin of this 
1489: `violation' of energy conservation, we can examine the form of the 
1490: singularity at $x^0 = x^0_c$, 
1491: $y=0$ (concentrating for definiteness on the $\tl<0$ case.) Defining
1492: \be \label{defth}
1493: \Theta \equiv e^{(x^0 - x^0_c)/\sqrt 2} = -e^{x^0/\sqrt 2}
1494: \sin(\tl\pi)\simeq 1
1495: \ee
1496: near the singularity, the singular part of $T_{00}$ near $(1-\Theta)\sim y
1497: \simeq 0$ is given 
1498: by:
1499: \be \label{esingpart}
1500: (T_{00})_{singular} = K(1 + \cos(\tl\pi)) \, {(1-\Theta) \over 
1501: (1-\Theta)^2 + {1\over 2} y^2} \simeq \, {\rm sgn}(1-\Theta) \, \pi
1502: \sqrt{2} \, K
1503: (1 + \cos(\tl\pi)) \, 
1504: \delta(y)\, .
1505: \ee
1506: Since $(1-\Theta)$ changes sign as $x^0$ passes from below to above
1507: $x^0_c$, there is a net loss of energy at $y=0$ at $x^0=x^0_c$,
1508: The amount of energy lost at $y=0$ as $x^0$ passes from below $x^0_c$
1509: to above $x^0_c$ is $2\pi \sqrt{2} \, K (1 + \cos(\tl\pi))$, -- same
1510: as the total 
1511: energy of the system. As a result
1512: the average of the energy 
1513: density in the rest of the system must vanish after the critical time
1514: $x^0_c$, as has already been seen in explicit calculation. If we
1515: continue the 
1516: formula for $T_{00}$ beyond this critical time, the system eventually
1517: evolves to a configuration with zero 
1518: energy density everywhere.\footnote{Of course, near the singularity
1519: $x^0=x^0_c$, $y=0$, the back reaction of closed string fields will be
1520: important, and a realistic analysis of the dynamics must include this
1521: effect.} There is an apparent violation of energy conservation at
1522: $x^0 = x^0_c$, but
1523: the most conservative interpretation of this is simply that we create a 
1524: codimension 1 D-brane at $y=0$, with a delta function source of energy
1525: that is not visible in the formula \refb{e5.20}. The total energy stored
1526: at $y=0$ does not 
1527: agree with the tension of a codimension 1 D-brane; in particular the ratio
1528: of the total energy of the system to that of a codimension 1 D-brane is
1529: given by:
1530: \be \label{eratio}
1531: {1\over \sqrt 2} (1 + \cos(\pi\tl))\, .
1532: \ee
1533: However the final state
1534: could represent a 
1535: state of the codimension 1 D-brane with non-trivial time dependent tachyon
1536: configuration, with the excess (deficit) energy stored in the 
1537: tachyon field. Fig. \ref{f2} shows the time evolution of $T_{00}$ given in
1538: eqs.\refb{e5.20}, \refb{e5.22} for a specific value of $\tl$.
1539: 
1540: 
1541: 
1542: \sectiono{Time Dependent Solutions on D-branes in Superstring 
1543: Theories} \label{s6}
1544: 
1545: The analysis of the previous sections can be easily generalized to the 
1546: case of superstring theories. Let us consider a generic D-brane system 
1547: with a tachyonic mode, and let the zero momentum tachyon vertex operator 
1548: have the form $e^{-\phi} V^{(-1)}_T$ in the $-1$ picture\cite{FMS}, where
1549: $\phi$ 
1550: denotes the bosonized ghost field and $V^{(-1)}_T$ denotes the matter part 
1551: of the vertex operator. If $V^{(-1)}_T$ has dimension $h<1/2$, then the 
1552: tachyon has mass$^2=(h-{1\over 2})\equiv -m^2$ and at the linearized 
1553: level the time dependent tachyon solution takes the form $\tl\cosh(m 
1554: x^0)$. In the Wick rotated theory this corresponds to $\tl\cos(mx)$, and 
1555: amounts to perturbing the BCFT by a term
1556: \be \label{e6.1}
1557: \tl \int dt\, d\theta \cos(m X(t,\theta)) V^{(-1)}_T(t,\theta)\, ,
1558: \ee
1559: where $t$ denotes the coordinate labelling the world-sheet boundary, 
1560: $\theta$ is the fermionic coordinate on the superspace, and we 
1561: consider $X$ and 
1562: $V_T^{(-1)}$ as 
1563: superfields. Although to lowest order in $\tl$ this represents a marginal 
1564: deformation, in general once higher order corrections are taken into 
1565: account this no longer represents a marginal deformation. We remedy this 
1566: by modifying \refb{e6.1} to 
1567: \be \label{e6.2}
1568: \tl \int dt \, d\theta \cos(\omg X(t,\theta)) V^{(-1)}_T(t,\theta)\, ,
1569: \ee
1570: where $\omg$ is a constant. The full $\beta$-function associated with 
1571: the coupling $\tl$ now takes the form:
1572: \be \label{e6.3}
1573: \omg_{\tl} = (\omg^2 - m^2)\tl + g(\omg,\tl)\, ,
1574: \ee
1575: as in \refb{e4.2}. We determine $\omg$ for a given $\tl$ by 
1576: demanding that $\beta_{\tl}$ vanishes. This gives:
1577: \be \label{e6.4}
1578: (\omg^2 - m^2)\tl + g(\omg,\tl) = 0\, .
1579: \ee
1580: This determines $\omg$ as a function of $\tl$. Thus for example the 
1581: leading correction to $\omg$ is given by
1582: \be \label{el1}
1583: \omg^2 \simeq m^2 - g(m, \tl) / \tl\, .
1584: \ee
1585: The $\beta$-functions associated with the other operators can be made to 
1586: vanish as in section \ref{s4} provided the product of \refb{e6.2} 
1587: with itself does not generate integral of any 
1588: superfield of dimension $1/2+\OO(\tl)$ other than the superfield 
1589: $\cos(\omg
1590: X(t,\theta)) V^{(-1)}_T(t,\theta)$ itself.
1591: 
1592: For an explicit example we can consider the case analogous to the one 
1593: discussed in section \ref{s5} for open bosonic string theory. We take a 
1594: non-BPS D-$p$-brane with one direction compactified on a circle of radius 
1595: $> {\sqrt 2}$. Let $y$ denote this compact direction. In this case the 
1596: first momentum mode of the tachyon along the circle, described by the 
1597: wave-function $\cos(y/R)$, can be thought of as a scalar field living
1598: in $((p-1)+1)$ 
1599: dimensions with mass$^2$:
1600: \be \label{e6.5}
1601: {1\over R^2} - {1\over 2} \equiv -m^2\, .
1602: \ee
1603: Displacing this tachyonic mode from its maximum and letting it roll 
1604: amounts to displacing the full $(p+1)$ dimensional tachyon from its 
1605: maximum by an amount proportional to $\cos(y/R)$ and let it roll. Since 
1606: the sign of the displacement is different for different values of $y$, we 
1607: expect that the time evolution of the tachyon will produce a kink-antikink 
1608: pair on the D-$p$-brane world volume, centered around $y=\pm\pi R/2$.
1609: 
1610: We shall study the time evolution by viewing this particular mode of the 
1611: tachyon as a scalar field in $((p-1)+1)$ dimensions.
1612: The vertex operator of this zero momentum tachyon in the $-1$ picture is 
1613: $e^{-\phi}\cos(Y/R)\otimes \sigma_1$
1614: where $\sigma_1$ is the Chan-Paton factor\cite{9808141,9809111}.
1615: Thus in order to construct the Wick rotated version of the rolling tachyon 
1616: solution, we need to perturb the BCFT by the operator
1617: \be \label{e6.6}
1618: \tl \int dt d\theta \cos(\omg X(t,\theta)) \cos(Y(t,\theta)/R)\otimes 
1619: \sigma_1\, .
1620: \ee
1621: If we expand the superfields $X(t,\theta)$ and $Y(t,\theta)$ as
1622: \be \label{e6.7}
1623: X(t,\theta) = X(t) + \theta\psi_x(t), \qquad
1624: Y(t,\theta) = Y(t) + \theta\psi_y(t),
1625: \ee
1626: where $\psi_x$ and $\psi_y$ are the world-sheet fermionic partners of the 
1627: $X$ and the $Y$ fields respectively, 
1628: then the perturbation can be expressed as
1629: \be \label{e6.8}
1630: -\tl \int dt [\omg \sin(\omg X(t)) \cos(Y(t)/R) + R^{-1} 
1631: \cos(\omg X(t)) \sin(Y(t)/R)]\otimes \sigma_1\, .
1632: \ee
1633: As in section \ref{s5}
1634: we can relate the computation of $g(m,\tl)$ appearing in 
1635: eq. \refb{el1} to leading order in $\tl$
1636: to the computation of the four point amplitude of four 
1637: on-shell external states, each described the the vertex operator 
1638: $e^{-\phi} \cos(m X) \cos(Y/R)\otimes \sigma_1$. This in turn will 
1639: determine the leading 
1640: correction to $\omg$ via eq.\refb{el1}. 
1641: We shall not carry out the detailed computation here. 
1642: 
1643: We can consider the special value of the radius $R=2$. In this case 
1644: $m=1/2$, and to leading order when we set $\omg=m$ the perturbation 
1645: \refb{e6.6} takes the form:
1646: \ben \label{e6.9}
1647: &&\tl \int dt d\theta \cos(X(t,\theta)/2) \cos(Y(t,\theta)/2)\otimes 
1648: \sigma_1 \nonumber \\
1649: &=& {\tl\over 2} \int dt \, d\theta \, [ \cos(Z_1(t,\theta)/\sqrt 2) + 
1650: \cos(Z_2(t,\theta)/\sqrt 2) 
1651: ]\otimes \sigma_1\, ,
1652: \een
1653: where
1654: \be \label{e6.10}
1655: Z_1 = (X+Y)/\sqrt 2, \qquad Z_2 = (X-Y)/\sqrt 2\, .
1656: \ee
1657: Writing
1658: \be \label{e6.11}
1659: Z_i(t,\theta) = Z_i(t) + \theta \psi_i(t)\, ,
1660: \ee
1661: we can rewrite the perturbation as
1662: \be \label{e6.12}
1663: -{\tl\over 2\sqrt 2} \int dt [\psi_1(t) \sin (Z_1(t)/\sqrt 2)
1664: + \psi_2(t) \sin (Z_2(t)/\sqrt 2)]\otimes \sigma_1\, .
1665: \ee
1666: Following the results of \cite{9808141,0003124} one can show that each of 
1667: the two terms 
1668: in \refb{e6.12} represents an exactly marginal deformation. Unfortunately 
1669: however 
1670: these two terms do not commute, instead they anti-commute due to the 
1671: presence of the fermion fields $\psi_i$. As a result the complete 
1672: perturbation is not exactly marginal and the $\beta$-function does receive 
1673: higher order corrections. Thus unlike in the case of section \ref{s5}, we 
1674: cannot solve the theory exactly in this case.
1675: In particular,
1676: we cannot obtain an analytic expression for the 
1677: boundary state and compute the energy momentum tensor associated with 
1678: this solution.
1679: 
1680: To see this more explicitly we fermionize  and rebosonize following the 
1681: procedure given in refs:\cite{9808141,0003124}. On the boundary the 
1682: relevant part of these 
1683: relations take the form:
1684: \be \label{e6.13}
1685: e^{iZ_i(t)/\sqrt 2} = {1\over \sqrt 2} (\xi^i(t) + i \eta^i(t)) \otimes 
1686: \tau_i\, ,
1687: \ee
1688: \be \label{e6.15}
1689: \psi_i(t) \eta_i(t) = -{i\over \sqrt 2} \p \phi_i(t)\, ,
1690: \ee
1691: where $\xi_i$ and $\eta_i$ are Majorana fermions, $\phi_i$ are scalars and 
1692: $\tau_i$ are Pauli matrices providing the cocycles needed for 
1693: the bosonization. In terms of the bosonic field $\phi_i$ the perturbation 
1694: \refb{e6.12} can be written as:
1695: \be \label{e6.16}
1696: {i\tl\over 4\sqrt 2} \int dt [\p_t\phi_1 \otimes \tau_1\otimes \sigma_1
1697: + \p_t\phi_2 \otimes \tau_2 \otimes \sigma_1]\, .
1698: \ee
1699: Since $\tau_1\otimes \sigma_1$ does not commute with 
1700: $\tau_2\otimes\sigma_1$, the above perturbation can be thought of as 
1701: switching on a pair of non-commuting Wilson lines, one along $\phi_1$ and 
1702: the other along $\phi_2$, each being of magnitude $\tl/4\sqrt 2$. As a 
1703: result the 
1704: potential does not vanish. Instead we get a contribution proportional to
1705: the square of the commutator, which in this case will be equal to $C\tl^4$ 
1706: for 
1707: an approprite positive constant $C$. This gives a contribution to 
1708: $g(m,\tl)$ equal 
1709: to $4 C\tl^3$. Eq.\refb{el1} then gives:
1710: \be \label{e6.17}
1711: \omg^2 = {1\over 4} - 4 C \tl^2\, .
1712: \ee
1713: This correction to the value of $\omg$ shows that \refb{e6.9} does not 
1714: represent an exactly marginal deformation.
1715: 
1716: If we had naively ignored the issue of non-commutativity of the two 
1717: operators in \refb{e6.16} and treated this as an exactly marginal 
1718: deformation, then we would end up with an expression for the energy 
1719: momentum tensor similar to that in \refb{e5.20}, with $f(x^0)$ given 
1720: by\cite{0203265}
1721: \be \label{e6.18}
1722: f(x^0) =  {1\over 1 + e^{\sqrt 2 x^0} \sin^2(\tl\pi/2)} + {1 \over
1723: 1 + e^{-\sqrt 2 x^0} \sin^2(\tl\pi/2)} - 1\, .
1724: \ee
1725: {}From this we see that $f((x^0\pm iy)/\sqrt{2})$ has singularities at 
1726: $x^0=\ln(1/\sin^2(\tl\pi/2))$, $y=\pm \pi$, and the energy density blows 
1727: up at these space-time points. Thus energy flows into these points from 
1728: the rest 
1729: of the system. These are precisely the points where we expect the 
1730: codimension 1 brane-antibrane pairs to form, and the nature of these 
1731: singularities has precisely the same delta function form as discussed at 
1732: the end of section \ref{s5}.\footnote{The analysis of RR charge density
1733: shows that RR charge also accumulates at these points.} Thus one might be
1734: tempted to interpret
1735: this as the creation of brane-antibrane pair.\footnote{Production of
1736: brane-antibrane pair from the decay of a higher dimensional brane has also
1737: been discussed in a different approximation in \cite{0204203}.} However,
1738: once we take into 
1739: account the non-marginality of the perturbation \refb{e6.16} this result 
1740: will be modified. In fact, if the perturbation \refb{e6.12} had been
1741: exactly marginal, and the energy momentum tensor had been given by
1742: \refb{e5.20}, \refb{e6.18}, we would have an inconsistency. To see this,
1743: note that following arguments similar to those in section \ref{s5} we
1744: would conclude in this case that the total initial energy of the system is
1745: deposited into a codimension 1 brane-antibrane pair, each carrying an
1746: energy density
1747: equal to:
1748: \be \label{e6.19}
1749: {1\over \sqrt 2} \, \TT_{p-1} \, (1 + \cos(\tl\pi))\, ,
1750: \ee
1751: where $\TT_{p-1}$ is the tension of a BPS D-$(p-1)$ brane. Thus for a
1752: generic $\tl$, \refb{e6.19} differs from the tension of a BPS
1753: D-$(p-1)$-brane. 
1754: Unlike in the case of bosonic D-branes, in this case we
1755: cannot attribute the excess (deficit) energy to the energy stored in the
1756: tachyon
1757: field, since there are no tachyons on a BPS D-brane. (The excess energy
1758: can still be attributed to other fields on the
1759: BPS D-brane, but there is no way a brane can have energy density lower
1760: than its tension.) Thus we would be led to an inconsistency, as there will
1761: be no interpretation for the final state for $(1+\cos(\pi\tl))< \sqrt{2}$.
1762: Thus
1763: it is just as well that \refb{e5.20}, \refb{e6.18} does not give the
1764: correct formula for the stress tensor.
1765: 
1766: \sectiono{Tachyon Dynamics in Closed String Theory} \label{s7}
1767: 
1768: In this section we briefly discuss possible extensions of the method 
1769: outlined in this paper to the study of tachyon dynamics in closed string 
1770: theory. 
1771: Given a zero 
1772: momentum
1773: tachyon field described by a dimension $(h,h)$ vertex operator $V_T$, we 
1774: try to construct a solution in the Wick rotated theory ($x^0\to i x$) by 
1775: perturbing the orginal CFT by 
1776: \be \label{e8.1}
1777: \tl \int d^2 z \, \cos(\omg X(z,\bar z)) V_T(z, \bar z)\, .
1778: \ee
1779: For $\omg^2 = 4(1-h)$ this would describe a marginal deformation to
1780: leading order in $\tl$. 
1781: However since in general higher order contribution to the $\beta$-function 
1782: does not vanish, we do not fix $\omg$ at the beginning. The 
1783: $\beta$-function associated with the perturbation \refb{e8.1} then takes 
1784: the form:
1785: \be \label{e8.2}
1786: \beta_{\tl} = 2\, \Big({1\over 4} \omg^2 + h - 1\Big)\, \tl + 
1787: g(\omg,\tl)\, .
1788: \ee
1789: We now adjust $\omg$ for a given $\tl$ such that \refb{e8.2} vanishes. 
1790: Thus this perturbation describes a conformal field theory.
1791: 
1792: The main obstruction to carrying out this program arises from the fact
1793: that
1794: in this case the operator product of \refb{e8.1} with itself does generate
1795: other dimension (1,1) operators, -- namely the vertex operators associated
1796: with the zero momentum graviton and dilaton fields.  (The source for the 
1797: dilaton field manifests itself in the form of a change in the central 
1798: charge induced by the perturbation.) Thus the rolling
1799: tachyon soution acts as a source for the zero momentum graviton and the
1800: dilaton fields, and {\it a priori} there is no guarantee that we shall be
1801: able to solve the equations of motion for these fields perturbatively in
1802: the parameter $\tl$. (See \cite{mukherji} for an analysis of this 
1803: problem in non-polynomial closed string field theory\cite{non-pol}.) This
1804: problem of course is not specific to string 
1805: theory, and even in the case of a normal scalar field theory coupled to 
1806: the 
1807: graviton and the dilaton, there is no guarantee that given a time 
1808: dependent solution involving the scalar field, we can solve for the 
1809: graviton and the dilaton field perturbatively in the parameter labelling 
1810: the solution of the scalar field equations. However, for tachyons 
1811: localized on a 
1812: subspace of the full space-time\cite{local}, it may be possible to solve 
1813: the 
1814: equations of motion of the graviton and the dilaton fields as a 
1815: perturbation expansion in the deformation parameter.
1816: 
1817: \sectiono{Conclusion} \label{s8}
1818: 
1819: In this paper we have discussed a general method for constructing time
1820: dependent solutions, describing the rolling of a tachyon on an unstable
1821: D-brane system, in cubic open string field theory. We have also provided a
1822: description of these solutions as Wick rotated version of euclidean
1823: boundary conformal field theories. The family of time dependent solutions
1824: associated with different initial conditions correspond to a family of
1825: boundary conformal field theories, each of which is related to the
1826: original D-brane system by a nearly marginal deformation. The construction
1827: can be easily generalized to the case of unstable D-brane systems in
1828: superstring theory.
1829: 
1830: In general these deformed boundary conformal field theories are not
1831: exactly solvable, and hence we cannot find explicit analytic expressions
1832: for the time dependence of the energy momentum tensor by Wick rotating the
1833: boundary state associated with the boundary conformal field theory. 
1834: However in some special cases the family of boundary CFT's are related by
1835: an exactly marginal deformation and are exactly solvable. This allows us
1836: to compute the time evolution of the energy momentum tensor explicitly for
1837: arbitrary initial condition on the tachyon. One such example discussed in
1838: this paper is the case of a bosonic D-$p$-brane wrapped on a circle of
1839: radius $\sqrt 2$, and we consider the first momentum mode of the tachyon
1840: along the circle. Pushing this particular tachyonic mode away from its
1841: extremum corresponds to an initial condition where the tachyon field has
1842: different signs in different parts of the circle, and hence we might
1843: expect that the time evolution of the configuration may produce lower
1844: dimensional D-branes. We explicitly construct the conserved energy
1845: momentum tensor associated with this solution and show that the time
1846: evolution of the system does indicate the creation of a codimension 1
1847: brane. 
1848: 
1849: One would like to generalize this to describe creation of a lower
1850: dimensional brane-antibrane pair due to the decay of an unstable D-brane
1851: in superstring theory.
1852: But unfortunately the situation here is more complicated as the associated
1853: family of conformal field theories are not related to each other by
1854: marginal deformation and are not exactly solvable. We hope to return to
1855: this problem in the near future.
1856: 
1857: We conclude with two remarks:
1858: 
1859: \begin{itemize}
1860: \item The solutions constructed here are obtained in {\it classical} open
1861: string field theory where we ignore coupling to closed string fields. This
1862: is expected to be a good approximation in the string coupling $g_s\to 0$
1863: limit. Once we switch on $g_s$, two effects will have to be taken into
1864: account in the study of the solution: classical backreaction of closed
1865: string fields on the solution, and the possibility of the decay into
1866: quantum closed string states. The time scale over which these effects will
1867: affect the solution is expected to go to infinity in the $g_s\to 0$ limit,
1868: but at present we do not know the precise dependence of this time scale on
1869: $g_s$.
1870: 
1871: Another parameter that is important in determining the effect of closed
1872: string coupling to the solution is $\tl$, or equivalently, the total
1873: energy
1874: density of the configuration. We would expect that for small energy
1875: density, the effect of coupling to the closed strings will be small, but
1876: again we do not know the precise dependence.
1877: 
1878: \item In the study of closed string tachyon condensation\cite{local},
1879: renormalization group (RG) flow has played a useful role. 
1880: The approach taken here is somewhat different: we interpret the complete
1881: time dependent solution for a specific initial condition as a single
1882: conformal field theory (after Wick rotation). Different inequivalent
1883: initial conditions will generate different boundary conformal field
1884: theories.
1885: 
1886: In fact it is possible to argue that at least for open string tachyon
1887: condensation, RG flow can never describe classical time
1888: evolution
1889: in a strict sense. RG flow takes us from a BCFT with higher boundary
1890: entropy to one of lower boundary entropy\cite{bent}. In the space-time
1891: language this corresponds to interpolating between a D-brane of higher
1892: energy to a D-brane of lower energy. It is clear that classical time
1893: evolution can never do that; due to conservation of energy it must take us
1894: from a D-brane of higher energy to one of lower energy plus other stuff.
1895: As a result, the RG analysis can at most give some qualitative information
1896: about time evolution in open string theory.
1897: 
1898: \end{itemize}
1899: 
1900: \medskip
1901: 
1902: {\bf Acknowledgement}:
1903: I would like to thank R.~Brower, R.~Gopakumar, J.~Minahan, S.~Minwalla, 
1904: N.~Moeller, L.~Rastelli and B.~Zwiebach for 
1905: useful discussions, and B.~Zwiebach for critical comments on the
1906: manuscript.
1907: This work was supported in part by a grant 
1908: from the Eberly College 
1909: of Science of the Penn State University. I would also like to acknowledge 
1910: the hospitality of the Center for Theoretical Physics at
1911: MIT, and a grant from the NM Rothschild and Sons Ltd
1912: at the Isaac Newton
1913: Institute where part of this work 
1914: was done.
1915: 
1916: 
1917: \begin{thebibliography}{99}
1918: 
1919: \bibitem{closed}
1920: G.~T.~Horowitz and A.~R.~Steif,
1921: Phys.\ Lett.\ B {\bf 258}, 91 (1991);
1922: %%CITATION = PHLTA,B258,91;%%
1923: G.~Veneziano,
1924: %``String cosmology: The pre-big bang scenario,''
1925: arXiv:hep-th/0002094;
1926: %%CITATION = HEP-TH 0002094;%%
1927: F.~Larsen and F.~Wilczek,
1928: Phys.\ Rev.\ D {\bf 55}, 4591 (1997)
1929: [arXiv:hep-th/9610252];
1930: %%CITATION = HEP-TH 9610252;%%
1931: G.~T.~Horowitz and D.~Marolf,
1932: JHEP {\bf 9807}, 014 (1998)
1933: [arXiv:hep-th/9805207];
1934: %%CITATION = HEP-TH 9805207;%%
1935: I.~I.~Kogan and N.~B.~Reis,
1936: Int.\ J.\ Mod.\ Phys.\ A {\bf 16}, 4567 (2001)
1937: [arXiv:hep-th/0107163];
1938: %%CITATION = HEP-TH 0107163;%%
1939: J.~Khoury, B.~A.~Ovrut, P.~J.~Steinhardt and N.~Turok,
1940: Phys.\ Rev.\ D {\bf 64}, 123522 (2001)
1941: [arXiv:hep-th/0103239];
1942: %%CITATION = HEP-TH 0103239;%%
1943: J.~Khoury, B.~A.~Ovrut, N.~Seiberg, P.~J.~Steinhardt and N.~Turok,
1944: Phys.\ Rev.\ D {\bf 65}, 086007 (2002)
1945: [arXiv:hep-th/0108187];
1946: %%CITATION = HEP-TH 0108187;%%
1947: P.~J.~Steinhardt and N.~Turok,
1948: arXiv:hep-th/0111030;
1949: %%CITATION = HEP-TH 0111030;%%
1950: N.~Seiberg,
1951: arXiv:hep-th/0201039;
1952: %%CITATION = HEP-TH 0201039;%%
1953: V.~Balasubramanian, S.~F.~Hassan, E.~Keski-Vakkuri and A.~Naqvi,
1954: %``A space-time orbifold: A toy model for a cosmological singularity,''
1955: arXiv:hep-th/0202187;
1956: %%CITATION = HEP-TH 0202187;%%
1957: L.~Cornalba and M.~S.~Costa,
1958: %``A New Cosmological Scenario in String Theory,''
1959: arXiv:hep-th/0203031;
1960: %%CITATION = HEP-TH 0203031;%%
1961: N.~A.~Nekrasov,
1962: %``Milne universe, tachyons, and quantum group,''
1963: arXiv:hep-th/0203112;
1964: %%CITATION = HEP-TH 0203112;%%
1965: J.~Simon,
1966: %``The geometry of null rotation identifications,''
1967: JHEP {\bf 0206}, 001 (2002)
1968: [arXiv:hep-th/0203201];
1969: %%CITATION = HEP-TH 0203201;%%
1970: E.~Kiritsis and B.~Pioline,
1971: arXiv:hep-th/0204004;
1972: %%CITATION = HEP-TH 0204004;%%
1973: H.~Liu, G.~Moore and N.~Seiberg,
1974: %``Strings in a time-dependent orbifold,''
1975: arXiv:hep-th/0204168;
1976: %%CITATION = HEP-TH 0204168;%%
1977: arXiv:hep-th/0206182;
1978: %%CITATION = HEP-TH 0206182;%%
1979: S.~Elitzur, A.~Giveon, D.~Kutasov and E.~Rabinovici,
1980: %``From big bang to big crunch and beyond,''
1981: arXiv:hep-th/0204189;
1982: %%CITATION = HEP-TH 0204189;%%
1983: L.~Cornalba, M.~S.~Costa and C.~Kounnas,
1984: arXiv:hep-th/0204261;
1985: %%CITATION = HEP-TH 0204261;%
1986: O.~Aharony, M.~Fabinger, G.~T.~Horowitz and E.~Silverstein, 
1987: %``Clean time-dependent string backgrounds from bubble baths,''
1988: arXiv:hep-th/0204158;
1989: %%CITATION = HEP-TH 0204158;%%
1990: B.~Craps, D.~Kutasov and G.~Rajesh,
1991: arXiv:hep-th/0205101;
1992: %%CITATION = HEP-TH 0205101;%%
1993: S.~Kachru, X.~Liu, M.~B.~Schulz and S.~P.~Trivedi,
1994: arXiv:hep-th/0205108;
1995: %%CITATION = HEP-TH 0205108;%%
1996: D.~Birmingham and M.~Rinaldi, 
1997: arXiv:hep-th/0205246;  
1998: %%CITATION = HEP-TH 0205246;%%
1999: A.~Lawrence,
2000: arXiv:hep-th/0205288;
2001: %%CITATION = HEP-TH 0205288;%% 
2002: V.~Balasubramanian and S.~F.~Ross, 
2003: %``The dual of nothing,''
2004: arXiv:hep-th/0205290;
2005: %%CITATION = HEP-TH 0205290;%%
2006: M.~Fabinger and J.~McGreevy,
2007: arXiv:hep-th/0206196;
2008: %%CITATION = HEP-TH 0206196;%%
2009: R.~Kallosh, L.~Kofman, A.~D.~Linde and A.~A.~Tseytlin,
2010: Phys.\ Rev.\ D {\bf 64}, 123524 (2001)
2011: [arXiv:hep-th/0106241];
2012: %%CITATION = HEP-TH 0106241;%%
2013: G.~T.~Horowitz and J.~Polchinski,
2014: arXiv:hep-th/0206228;
2015: %%CITATION = HEP-TH 0206228;%%
2016: C.~M.~Chen, D.~V.~Gal'tsov and M.~Gutperle,
2017: arXiv:hep-th/0204071;
2018: %%CITATION = HEP-TH 0204071;%%
2019: M.~Kruczenski, R.~C.~Myers and A.~W.~Peet,
2020: JHEP {\bf 0205}, 039 (2002)
2021: [arXiv:hep-th/0204144];
2022: %%CITATION = HEP-TH 0204144;%%
2023: S.~Roy,
2024: arXiv:hep-th/0205198;
2025: %%CITATION = HEP-TH 0205198;%%
2026: N.~S.~Deger and A.~Kaya,
2027: arXiv:hep-th/0206057;
2028: %%CITATION = HEP-TH 0206057;%%
2029: J.~E.~Wang,
2030: arXiv:hep-th/0207089.
2031: %%CITATION = HEP-TH 0207089;%%
2032: 
2033: \bibitem{0202210}
2034: M.~Gutperle and A.~Strominger,
2035: %``Spacelike branes,''
2036: arXiv:hep-th/0202210.
2037: %%CITATION = HEP-TH 0202210;%%
2038: 
2039: 
2040: 
2041: \bibitem{0203211} 
2042: A.~Sen, 
2043: %``Rolling Tachyon,'' 
2044: arXiv:hep-th/0203211. 
2045: %%CITATION = HEP-TH 0203211;%%
2046: 
2047: \bibitem{0203265}
2048: A.~Sen,
2049: %``Tachyon matter,''
2050: arXiv:hep-th/0203265.
2051: %%CITATION = HEP-TH 0203265;%%
2052: 
2053: \bibitem{0204203}
2054: K.~Hashimoto,
2055: arXiv:hep-th/0204203.
2056: %%CITATION = HEP-TH 0204203;%%
2057: 
2058: \bibitem{0205085}
2059: S. Sugimoto, S. Terashima
2060: arXiv:hep-th/0205085.
2061: %%CITATION = HEP-TH 0205085;%%
2062: 
2063: \bibitem{0205098}
2064: J.~A.~Minahan,
2065: %``Rolling the tachyon in super BSFT,''
2066: arXiv:hep-th/0205098.
2067: %%CITATION = HEP-TH 0205098;%%
2068: 
2069: \bibitem{0206102}
2070: A.~Ishida and S.~Uehara,
2071: %``Gauge fields on tachyon matter,''
2072: arXiv:hep-th/0206102.
2073: %%CITATION = HEP-TH 0206102;%%
2074: 
2075: \bibitem{0207004}
2076: K.~Ohta and T.~Yokono, 
2077: arXiv:hep-th/0207004
2078: %%CITATION = HEP-TH 0207004;%%
2079: 
2080: \bibitem{old}
2081: K.~Bardakci, 
2082: Nucl.\ Phys.\ B {\bf 68}, 331 (1974)
2083: %%CITATION = NUPHA,B68,331;%%
2084: Nucl.\ Phys.\ B {\bf 70}, 397 (1974);
2085: %%CITATION = NUPHA,B70,397;%%
2086: Nucl.\ Phys.\ B {\bf 133}, 297 (1978);
2087: %%CITATION = NUPHA,B133,297;%%
2088: K.~Bardakci and M.~B.~Halpern,
2089: Phys.\ Rev.\ D {\bf 10}, 4230 (1974);
2090: %%CITATION = PHRVA,D10,4230;%%
2091: Nucl.\ Phys.\ B {\bf 96}, 285 (1975).
2092: %%CITATION = NUPHA,B96,285;%%
2093: 
2094: \bibitem{9806155}
2095: O.~Bergman and M.~R.~Gaberdiel,
2096: %``Stable non-BPS D-particles,''
2097: Phys.\ Lett.\ B {\bf 441}, 133 (1998)
2098: [arXiv:hep-th/9806155];
2099: %%CITATION = HEP-TH 9806155;%%
2100: %``Non-BPS Dirichlet branes,''
2101: Class.\ Quant.\ Grav.\  {\bf 17}, 961 (2000)
2102: [arXiv:hep-th/9908126].
2103: %%CITATION = HEP-TH 9908126;%%
2104: 
2105: 
2106: 
2107: \bibitem{9809111}
2108: A.~Sen,
2109: %``Type I D-particle and its interactions,''
2110: JHEP {\bf 9810}, 021 (1998)
2111: [arXiv:hep-th/9809111].
2112: %%CITATION = HEP-TH 9809111;%%
2113: 
2114: 
2115: \bibitem{cosmo}
2116: G.~W.~Gibbons,
2117: %``Cosmological evolution of the rolling tachyon,''
2118: arXiv:hep-th/0204008;
2119: %%CITATION = HEP-TH 0204008;%%
2120: M.~Fairbairn and M.~H.~Tytgat,
2121: %``Inflation from a tachyon fluid?,''
2122: arXiv:hep-th/0204070;
2123: %%CITATION = HEP-TH 0204070;%%
2124: S.~Mukohyama,
2125: %``Brane cosmology driven by the rolling tachyon,''
2126: arXiv:hep-th/0204084;
2127: %%CITATION = HEP-TH 0204084;%%
2128: A.~Feinstein,
2129: %``Power-law inflation from the rolling tachyon,''
2130: arXiv:hep-th/0204140;
2131: %%CITATION = HEP-TH 0204140;%%
2132: T.~Padmanabhan, arXiv:hep-th/0204150;
2133: %%CITATION = HEP-TH 0204150;%%
2134: A.~Frolov, L.~Kofman and A.~A.~Starobinsky,
2135: arXiv:hep-th/0204187;
2136: %%CITATION = HEP-TH 0204187;%%
2137: D.~Choudhury, D.~Ghoshal, D.~P.~Jatkar and S.~Panda,
2138: arXiv:hep-th/0204204;
2139: %%CITATION = HEP-TH 0204204;%%
2140: X.~Li, J.~Hao and D.~Liu,
2141: arXiv:hep-th/0204252;
2142: %%CITATION = HEP-TH 0204252;%%
2143: G.~Shiu and I.~Wasserman,
2144: arXiv:hep-th/0205003;
2145: %%CITATION = HEP-TH 0205003;%%
2146: T.~Padmanabhan and T.~R.~Choudhury,
2147: %``Can the clustered dark matter and the smooth dark energy arise from the  
2148: %same scalar field?,''
2149: arXiv:hep-th/0205055;
2150: %%CITATION = HEP-TH 0205055;%%
2151: L.~Kofman and A.~Linde,
2152: %``Problems with Tachyon Inflation,''
2153: arXiv:hep-th/0205121;
2154: %%CITATION = HEP-TH 0205121;%%
2155: M.~Sami,
2156: %``Implementing Power Law Inflation with Rolling Tachyon on the Brane,''
2157: arXiv:hep-th/0205146;
2158: %%CITATION = HEP-TH 0205146;%%
2159: M.~Sami, P.~Chingangbam and T.~Qureshi,
2160: %``Aspects of tachyonic inflation with exponential potential,''
2161: arXiv:hep-th/0205179;
2162: %%CITATION = HEP-TH 0205179;%%
2163: T.~Mehen and B.~Wecht,
2164: arXiv:hep-th/0206212.
2165: %%CITATION = HEP-TH 0206212;%%
2166: 
2167: \bibitem{earl}
2168: G.~R.~Dvali and S.~H.~Tye, Phys. \ Lett. \ B {\bf 450}, 72 (1999)
2169: [arXiv:hep-ph/9812483];
2170: %%CITATION = HEP-TH 9812483;%%
2171: C.~Acatrinei and C.~Sochichiu,
2172: %``A note on the decay of noncommutative solitons,''
2173: arXiv:hep-th/0104263;
2174: %%CITATION = HEP-TH 0104263;%%
2175: S.~H.~Alexander, Phys. \ Rev. \ D {\bf 65}, 023507 (2002)
2176: [arXiv:hep-th/0105032];
2177: %%CITATION = HEP-TH 0105032;%%
2178: G.~R.~Dvali, Q.~Shafi and S.~Solganik,
2179: [arXiv:hep-th/0105203];
2180: %%CITATION = HEP-TH 0105203;%%
2181: C.~P.~Burgess, M.~Majumdar, D.~Nolte, F.~Quevedo, G.~Rajesh and 
2182: R.~Z.~Zhang, JHEP {\bf 0107}, 047 (2001)
2183: [arXiv:hep-th/0105204];
2184: %%CITATION = HEP-TH 0105204;%%
2185: A.~Mazumdar, S.~Panda and A.~Perez-Lorenzana,
2186: %``Assisted inflation via tachyon condensation,''
2187: Nucl.\ Phys.\ B {\bf 614}, 101 (2001)
2188: [arXiv:hep-ph/0107058];
2189: %%CITATION = HEP-PH 0107058;%%
2190: C.~P.~Burgess, P.~Martineau, F.~Quevedo, G.~Rajesh, and R.~J.~Zhang,
2191: JHEP {\bf 0203}, 052 (2002)
2192: [arXiv:hep-th/0111025];
2193: %%CITATION = HEP-TH 0111025;%%
2194: S.~Sarangi and S.~H.~Tye, 
2195: %``Cosmic string production towards the end of brane inflation,''
2196: arXiv:hep-th/0204074;
2197: %%CITATION = HEP-TH 0204074;%%
2198: S.~Corley and D.~A.~Lowe,
2199: Phys.\ Lett.\ B {\bf 528}, 139 (2002)
2200: [arXiv:hep-th/0108178].
2201: %%CITATION = HEP-TH 0108178;%%
2202: 
2203: \bibitem{9704006}
2204: E.~Gava, K.~S.~Narain and M.~H.~Sarmadi,
2205: %``On the bound states of p- and (p+2)-branes,''
2206: Nucl.\ Phys.\ B {\bf 504}, 214 (1997)
2207: [arXiv:hep-th/9704006].
2208: %%CITATION = HEP-TH 9704006;%%
2209: 
2210: \bibitem{9703217}
2211: A.~Hashimoto and W.~I.~Taylor,
2212: %``Fluctuation spectra of tilted and intersecting D-branes from the  
2213: %Born-Infeld action,''
2214: Nucl.\ Phys.\ B {\bf 503}, 193 (1997)
2215: [arXiv:hep-th/9703217].
2216: %%CITATION = HEP-TH 9703217;%%
2217: 
2218: \bibitem{inter}
2219: J.~Garcia-Bellido, R.~Rabadan and F.~Zamora,
2220: %``Inflationary scenarios from branes at angles,''
2221: JHEP {\bf 0201}, 036 (2002)
2222: [arXiv:hep-th/0112147];
2223: %%CITATION = HEP-TH 0112147;%%
2224: R.~Blumenhagen, B.~Kors, D.~Lust and T.~Ott,
2225: %``Hybrid inflation in intersecting brane worlds,''
2226: arXiv:hep-th/0202124;
2227: %%CITATION = HEP-TH 0202124;%%
2228: K.~Dasgupta, C.~Herdeiro, S.~Hirano and R.~Kallosh,
2229: %``D3/D7 inflationary model and M-theory,''
2230: Phys.\ Rev.\ D {\bf 65}, 126002 (2002)
2231: [arXiv:hep-th/0203019];
2232: %%CITATION = HEP-TH 0203019;%%
2233: N.~Jones, H.~Stoica and S.~H.~Tye,
2234: %``Brane interaction as the origin of inflation,''
2235: arXiv:hep-th/0203163.
2236: %%CITATION = HEP-TH 0203163;%%
2237: 
2238: \bibitem{mo-zw}
2239: N.~Moeller and B.~Zwiebach, 
2240: arXiv:hep-th/0207107.
2241: %%CITATION = HEP-TH 0207107;%%
2242: 
2243: 
2244: \bibitem{golds}
2245: H.~Goldstein, Classical Mechanics, Addison Wesley.
2246: 
2247: \bibitem{zam}
2248: A.~B.~Zamolodchikov,
2249: %``Renormalization Group And Perturbation Theory Near Fixed Points In 
2250: %Two-Dimensional Field Theory,''
2251: Sov.\ J.\ Nucl.\ Phys.\  {\bf 46}, 1090 (1987)
2252: [Yad.\ Fiz.\  {\bf 46}, 1819 (1987)].
2253: %%CITATION = SJNCA,46,1090;%%
2254: 
2255: \bibitem{bent} 
2256: I.~Affleck and A.~W.~Ludwig, 
2257: Phys.\ Rev.\ Lett.\ {\bf 67}, 161 (1991);  
2258: %%CITATION = PRLTA,67,161;%%,
2259: P.~Fendley, H.~Saleur and N.~P.~Warner,
2260: Nucl.\ Phys.\ B {\bf 430}, 577 (1994)
2261: [arXiv:hep-th/9406125];
2262: %%CITATION = HEP-TH 9406125;%%
2263: J.~A.~Harvey, D.~Kutasov and E.~J.~Martinec, 
2264: arXiv:hep-th/0003101.  
2265: %%CITATION = HEP-TH 0003101;%%
2266: 
2267: 
2268: \bibitem{bsft}
2269: A.~A.~Gerasimov and S.~L.~Shatashvili,
2270: %``On exact tachyon potential in open string field theory,''
2271: JHEP {\bf 0010}, 034 (2000)
2272: [arXiv:hep-th/0009103];
2273: %%CITATION = HEP-TH 0009103;%%
2274: D.~Kutasov, M.~Marino and G.~W.~Moore,
2275: %``Some exact results on tachyon condensation in string field theory,''
2276: JHEP {\bf 0010}, 045 (2000)
2277: [arXiv:hep-th/0009148];
2278: %%CITATION = HEP-TH 0009148;%%
2279: %``Remarks on tachyon condensation in superstring field theory,''
2280: arXiv:hep-th/0010108;
2281: %%CITATION = HEP-TH 0010108;%%
2282: D.~Ghoshal and A.~Sen,
2283: JHEP {\bf 0011}, 021 (2000)
2284: [arXiv:hep-th/0009191].
2285: %%CITATION = HEP-TH 0009191;%%
2286: 
2287: \bibitem{0010247}
2288: S.~Dasgupta and T.~Dasgupta,
2289: JHEP {\bf 0106}, 007 (2001)
2290: [arXiv:hep-th/0010247].
2291: %%CITATION = HEP-TH 0010247;%%
2292: 
2293: 
2294: \bibitem{boundary} 
2295: J.~Polchinski and Y.~Cai, 
2296: %``Consistency Of Open Superstring Theories,'' 
2297: Nucl.\ Phys.\ B {\bf 296}, 91 (1988); 
2298: %%CITATION = NUPHA,B296,91;%%
2299: C.~G.~Callan, C.~Lovelace, C.~R.~Nappi and S.~A.~Yost,
2300: Nucl.\ Phys.\ B {\bf 460}, 351 (1996)[arXiv:hep-th/9510161];
2301: %%CITATION = HEP-TH 9510161;%%
2302: M.~Li,
2303: Nucl.\ Phys.\ B {\bf 460}, 351 (1996)
2304: [arXiv:hep-th/9510161];
2305: %%CITATION = HEP-TH 9510161;%%
2306: O.~Bergman and M.~R.~Gaberdiel,
2307: Nucl.\ Phys.\ B {\bf 499}, 183 (1997)
2308: [arXiv:hep-th/9701137].
2309: %%CITATION = HEP-TH 9701137;%%
2310: 
2311: \bibitem{em-tensor}
2312: M.~B.~Green and M.~Gutperle,
2313: Nucl.\ Phys.\ B {\bf 476}, 484 (1996)
2314: [arXiv:hep-th/9604091];
2315: %%CITATION = HEP-TH 9604091;%%
2316: P.~Di Vecchia, M.~Frau, I.~Pesando, S.~Sciuto, A.~Lerda and R.~Russo,
2317: Nucl.\ Phys.\ B {\bf 507}, 259 (1997)
2318: [arXiv:hep-th/9707068];
2319: %%CITATION = HEP-TH 9707068;%%
2320: P.~Di Vecchia and A.~Liccardo,
2321: arXiv:hep-th/9912275.
2322: %%CITATION = HEP-TH 9912275;%%
2323: 
2324: 
2325: 
2326: \bibitem{marginal}
2327: C.~G.~Callan and I.~R.~Klebanov,
2328: Phys.\ Rev.\ Lett.\  {\bf 72}, 1968 (1994)
2329: [arXiv:hep-th/9311092];
2330: %%CITATION = HEP-TH 9311092;%%
2331: C.G.~Callan, I.R.~Klebanov, A.W.~Ludwig and J.M.~Maldacena,
2332: Nucl.\ Phys.\ {\bf B422}, 417 (1994)
2333: hep-th/9402113;
2334: J.~Polchinski and L.~Thorlacius,
2335: Phys.\ Rev.\ {\bf D50}, 622 (1994)
2336: hep-th/9404008.
2337: %%CITATION = PHRVA,D50,622;%%
2338: %%CITATION = NUPHA,B422,417;%%
2339: 
2340: \bibitem{9811237}
2341: A.~Recknagel and V.~Schomerus, 
2342: %``Boundary deformation theory and moduli
2343: %spaces of D-branes,'',
2344: Nucl.\ Phys.\ {\bf B545}, 233 (1999)
2345: hep-th/9811237.  
2346: %%CITATION = NUPHA,B545,233;%%
2347: 
2348: \bibitem{0108238}
2349: M.~R.~Gaberdiel and A.~Recknagel,
2350: %``Conformal boundary states for free bosons and fermions,''
2351: JHEP {\bf 0111}, 016 (2001)
2352: [arXiv:hep-th/0108238].
2353: %%CITATION = HEP-TH 0108238;%%
2354: 
2355: \bibitem{FMS}
2356: D.~Friedan, E.~J.~Martinec and S.~H.~Shenker,
2357: Nucl.\ Phys.\ B {\bf 271}, 93 (1986).
2358: %%CITATION = NUPHA,B271,93;%%
2359: 
2360: \bibitem{9808141}
2361: A.~Sen,
2362: JHEP {\bf 9809}, 023 (1998)
2363: [hep-th/9808141]; JHEP {\bf 9812}, 021 (1998)
2364: [arXiv:hep-th/9812031].
2365: %%CITATION = HEP-TH 9812031;%%
2366: %%CITATION = HEP-TH 9808141;%%
2367: 
2368: 
2369: \bibitem{0003124}
2370: J.~Majumder and A.~Sen,  
2371: JHEP {\bf 0006}, 010 (2000) 
2372: [arXiv:hep-th/0003124].  
2373: %%CITATION = HEP-TH 0003124;%%
2374: 
2375: 
2376: \bibitem{mukherji}
2377: S.~Mukherji and A.~Sen,
2378: %``Some all order classical solutions in nonpolynomial closed string field 
2379: %theory,''
2380: Nucl.\ Phys.\ B {\bf 363}, 639 (1991).
2381: %%CITATION = NUPHA,B363,639;%%
2382: 
2383: \bibitem{non-pol}
2384: B.~Zwiebach,
2385: Nucl.\ Phys.\ B {\bf 390}, 33 (1993)
2386: [arXiv:hep-th/9206084];
2387: %%CITATION = HEP-TH 9206084;%%
2388: arXiv:hep-th/9305026;  
2389: %%CITATION = HEP-TH 9305026;%% 
2390: Annals Phys.\ {\bf 267}, 193 (1998)  
2391: [arXiv:hep-th/9705241].  
2392: %%CITATION = HEP-TH 9705241;%%
2393: 
2394: \bibitem{local}
2395: A.~Adams, J.~Polchinski and E.~Silverstein,
2396: %``Don't panic! Closed string tachyons in ALE space-times,''
2397: JHEP {\bf 0110}, 029 (2001)
2398: [arXiv:hep-th/0108075];
2399: %%CITATION = HEP-TH 0108075;%%
2400: A.~Dabholkar,
2401: %``On condensation of closed-string tachyons,''
2402: arXiv:hep-th/0109019;
2403: %%CITATION = HEP-TH 0109019;%%
2404: J.~G.~Russo and A.~A.~Tseytlin,
2405: %``Supersymmetric fluxbrane intersections and closed string tachyons,''
2406: JHEP {\bf 0111}, 065 (2001)
2407: [arXiv:hep-th/0110107];
2408: %%CITATION = HEP-TH 0110107;%%
2409: A.~Dabholkar,
2410: %``Tachyon condensation and black hole entropy,''
2411: Phys.\ Rev.\ Lett.\  {\bf 88}, 091301 (2002)
2412: [arXiv:hep-th/0111004];
2413: %%CITATION = HEP-TH 0111004;%%
2414: C.~Vafa,
2415: %``Mirror symmetry and closed string tachyon condensation,''
2416: arXiv:hep-th/0111051;
2417: %%CITATION = HEP-TH 0111051;%%
2418: J.~A.~Harvey, D.~Kutasov, E.~J.~Martinec and G.~Moore,
2419: %``Localized tachyons and RG flows,''
2420: arXiv:hep-th/0111154;
2421: %%CITATION = HEP-TH 0111154;%%
2422: A.~Dabholkar and C.~Vafa,
2423: %``tt* geometry and closed string tachyon potential,''
2424: JHEP {\bf 0202}, 008 (2002)
2425: [arXiv:hep-th/0111155];
2426: %%CITATION = HEP-TH 0111155;%%
2427: Y.~Michishita and P.~Yi,
2428: %``D-brane probe and closed string tachyons,''
2429: Phys.\ Rev.\ D {\bf 65}, 086006 (2002)
2430: [arXiv:hep-th/0111199];
2431: %%CITATION = HEP-TH 0111199;%%
2432: J.~R.~David, M.~Gutperle, M.~Headrick and S.~Minwalla,
2433: %``Closed string tachyon condensation on twisted circles,''
2434: JHEP {\bf 0202}, 041 (2002)
2435: [arXiv:hep-th/0111212];
2436: %%CITATION = HEP-TH 0111212;%%
2437: S.~k.~Nam and S.~J.~Sin,
2438: %``Condensation of localized tachyons and spacetime supersymmetry,''
2439: arXiv:hep-th/0201132;
2440: %%CITATION = HEP-TH 0201132;%%
2441: S.~P.~De Alwis and A.~T.~Flournoy,
2442: %``Closed string tachyons and semi-classical instabilities,''
2443: arXiv:hep-th/0201185;
2444: %%CITATION = HEP-TH 0201185;%%
2445: S.~J.~Sin,
2446: %``Tachyon mass, c-function and counting localized degrees of freedom,''
2447: arXiv:hep-th/0202097;
2448: %%CITATION = HEP-TH 0202097;%%
2449: R.~Rabadan and J.~Simon,
2450: %``M-theory lift of brane-antibrane systems and localised closed string  
2451: tachyons,''
2452: JHEP {\bf 0205}, 045 (2002)
2453: [arXiv:hep-th/0203243];
2454: %%CITATION = HEP-TH 0203243;%%
2455: A.~Basu,
2456: %``Localized tachyons and the g(cl) conjecture,''
2457: arXiv:hep-th/0204247.
2458: %%CITATION = HEP-TH 0204247;%%
2459: 
2460: \end{thebibliography}
2461: 
2462: \end{document}
2463: 
2464: \bye
2465: 
2466: