hep-th0403170/v2.tex
1: %--------------------------------------------------------------------+
2: % version 2, tue-20-mar-2004 @16:00 edt
3: %--------------------------------------------------------------------+
4: \documentclass[12pt]{article}
5: \textheight=9.0truein
6: \textwidth=6.5truein
7: \voffset=-0.65truein
8: \hoffset=-0.6truein
9: \overfullrule=0pt
10: \parskip=0pt
11: \baselineskip=10pt
12: \usepackage{amssymb}
13: \usepackage{epsfig}
14: \usepackage{amsmath}
15: %--------------------------------------------------------------------+
16: \begin{document}
17: 
18: \rightline{hep-th/0403170}
19: 
20: \bigskip\bigskip
21: 
22: \begin{center} 
23: {\Large \bf Brane-antibrane systems and \\ the thermal life of neutral
24: black holes}
25: \end{center} 
26: 
27: \bigskip\bigskip
28:   
29: \renewcommand{\thefootnote}{\fnsymbol{footnote}} 
30: \centerline{\bf
31: Omid~Saremi\footnote{omidsar@physics.utoronto.ca} and
32: Amanda~W.~Peet\footnote{peet@physics.utoronto.ca}}
33: \bigskip
34: \centerline{\it Department of Physics,}
35: \centerline{\it University of Toronto,}
36: \centerline{\it 60 St. George Street,}
37: \centerline{\it Toronto, Ontario,}
38: \centerline{\it Canada M5S 1A7.}
39:    
40: \setcounter{footnote}{0}   
41: \renewcommand{\thefootnote}{\arabic{footnote}}   
42: 
43: \bigskip\bigskip
44: 
45: \abstract{A brane-antibrane model for the entropy of neutral black
46: branes is developed, following on from the work of Danielsson, Guijosa
47: and Kruczenski [1].  The model involves equal numbers of D$p$-branes
48: and anti-D$p$-branes, and arbitrary angular momenta, and covers the
49: cases $p=0,1,2,3,4$.  The thermodynamic entropy is reproduced by the
50: strongly coupled field theory, up to a power of two.  The
51: strong-coupling physics of the $p=0$ case is further developed
52: numerically, using techniques of Kabat, Lifschytz et al. [2,3], in the
53: context of a toy model containing the tachyon and the bosonic degrees
54: of freedom of the D0-brane and anti-D0-brane quantum mechanics.
55: Preliminary numerical results show that strong-coupling
56: finite-temperature stabilization of the tachyon is possible, in this
57: context.}
58: 
59: \vfill
60: 
61: \noindent 16 March 2004.
62: 
63: \setcounter{page}{0}
64: %====================================================================+
65: \newpage
66: \section{Introduction}
67: %
68: The drive to explain the thermodynamic entropy of black holes and
69: black branes via the statistical mechanics of microscopic degrees of
70: freedom has been a central preoccupation of string theorists and other
71: gravitational theorists for at least two and a half decades.  String
72: theory has provided significant progress on this front in the last
73: eight years.  The celebrated 1996 success of Strominger and Vafa in
74: computing the entropy of particular $D=5$ supersymmetric (BPS) black
75: holes occurring in low-energy string theory, by using a conformal
76: field theory appropriate to the microscopic physics of strings and
77: D-branes, even raised the profile of string theory itself in a
78: significant way.  Further successes followed; the thermodynamic
79: entropy of BPS and near-BPS black holes in various dimensions was
80: reproduced successfully, in some cases even beyond leading order in
81: macroscopic quantum numbers such as mass, charge, and angular momenta.
82: A microscopic string theoretic accounting of the entropy of neutral
83: black holes and black branes remains more elusive, however. One reason
84: is that neutral black holes and branes are literally as far from BPS
85: as possible, meaning that supersymmetry may provide no help at all in
86: the endeavour.
87: 
88: BFSS matrix theory \cite{BFSS} is a conjectured relationship between
89: quantum $D=11$ M theory in light-front frame and the supersymmetric
90: quantum mechanics of a large number $N$ of D0-branes.  This theory was
91: recruited, e.g. in \cite{KlebanovSusskind}, to help explain the
92: entropy of near-BPS black holes in dimensions $D=10-p$, and also in
93: attempts to explain the entropy of the neutral cases as well.  Reviews
94: of this story include \cite{peetCQGreview} (see references therein).
95: Another interesting proposal was in \cite{Englert}.  Here, we simply
96: record one major common feature of these various matrix theory
97: approaches to counting the entropy of neutral black holes, which is
98: relevant to our work here.  Namely, the need for an infinite boost in
99: the 11th dimension $x^{11}$ in order to relate the neutral black holes
100: of interest to nearly-BPS systems whose entropy can be computed
101: microscopically using D-brane field theory.  In particular, this
102: infinite boost in the 11th dimension eliminates anti-D0-branes from
103: the picture.  The approach that we will take here will be different.
104: Other interesting approaches to neutral black hole entropy from quite
105: different perspectives include \cite{dasetal} and references therein.
106: 
107: Gravity/gauge correspondences are specific relationships between open
108: string and closed string degrees of freedom.  For the case of sixteen
109: supercharges \cite{itzhaki}, they relate the degrees of freedom of
110: supersymmetric Yang-Mills theories to closed string theory on
111: near-horizon D$p$-brane spacetimes.  Subsequent string theoretic
112: developments of interest for our work here include tachyon
113: condensation in brane-antibrane systems.  The paper which sparked our
114: specific interest was \cite{Danielsson}, in which an interesting
115: proposal was made for a microscopic explanation of the entropy of
116: neutral black D3, M2 and M5 (non-dilatonic) branes in terms of a
117: system of strongly coupled branes and antibranes.
118: 
119: In section 2, we review features of earlier analytic work, upon which
120: we build here.  Section 3 contains the bulk of our analytic
121: observations.  We study the entropy of neutral black D$p$-brane
122: spacetimes by using a microscopic model with an equal number of
123: strongly coupled D$p$-branes and anti-D$p$-branes, with arbitrary
124: angular momenta turned on.  Our main result is that the supergravity
125: entropy density can be reproduced by strongly coupled field theory, up
126: to a power of two.  Section 4 contains an exposition of preliminary
127: numerical work on the $p=0$ system.  Our primary goal in the numerical
128: part of our work is to try to test some assumptions of the
129: brane-antibrane model, by doing direct strong-coupling simulation of
130: the $p=0$ system, adapting methods of \cite{kabat1,kabat2}.  We use a
131: bosonic toy model to do this analysis, and find preliminary evidence
132: that tachyon stabilization may indeed occur in the fashion expected
133: from the work of \cite{Danielsson}.  We are not yet able, however, to
134: be conclusive about the supersymmetric case.
135: 
136: While this paper was in preparation, the work of \cite{Guijosa}
137: appeared, which has some overlap with section 3, for the case $p=3$
138: with angular momentum.
139: 
140: %====================================================================+
141: \section{Analytic ingredients from prior work}
142: 
143: %--------------------------------------------------------------------+
144: \subsection{Gravity/gauge correspondences for general $p$}
145: 
146: For a system of $N$ D$p$-branes, it is possible\footnote{For $p\leq
147: 4$, which are the cases on which we will concentrate} to take a clean
148: low-energy limit such that the open strings with endpoints on the
149: D$p$-branes decouple from the closed strings in the bulk.  This
150: observation led, of course, to the celebrated AdS/CFT correspondence
151: and the non-conformal open/closed-string correspondences of Itzhaki,
152: Maldacena, Sonnenschein and Yankielowicz \cite{itzhaki}.  The theory
153: on the branes is supersymmetric Yang-Mills theory (SYM) with sixteen
154: supercharges.  The 't Hooft coupling for the SYM theory is
155: %
156: \begin{equation}
157: g_{\rm YM}^2N\equiv(2\pi)^{p-2}g_s\ell_s^{p-3}\ N \,,
158: \end{equation}
159: %
160: where $g_s$ is the string coupling and
161: $\ell_s\equiv\sqrt{\alpha^\prime}$ is the string length.
162: 
163: These open/closed string correspondences tell us that we can hope to
164: understand the gravitational fields of D$p$-branes via the dual SYM
165: theory.  For $p\not=3$ the fields describing the $D=10$ geometry vary
166: radially in the spacetime, and on the SYM side the coupling is
167: dimensionful, yielding breakdown of the weak-coupling description
168: either in the UV or the IR.  In other words, there are limits to the
169: validity of both descriptions.  A dimensionless control variable is
170: given by $g_{\rm eff}^2\equiv g_{\rm YM}^2N U^{p-3}$; on the
171: supergravity side, $U\equiv r/\ell_s^2$ is the radial isotropic
172: coordinate in energy units.  The requirements \cite{itzhaki} for the
173: supergravity geometry in $D=10$ to remain valid are
174: %
175: \begin{equation}
176: 1\ll (g_{\rm YM}^2N U^{p-3})  \ll N^{\frac{4}{(7-p)}} \,.
177: \end{equation}
178: %
179: At the left-hand end, $\alpha'$ corrections become important and the
180: SYM theory takes over from the $D=10$ geometry as the weakly-coupled
181: description; at the right-hand end, strong coupling (dilaton) ensues
182: and it is necessary (for large-$N$) to turn to the S-dual supergravity
183: geometry.
184: 
185: It was earlier noted in \cite{KlebanovTseytlin} that thermodynamic
186: properties of near-BPS D$p$-brane (and M-brane) geometries could be
187: written in a way that is suggestive of a field theory interpretation.
188: The main focus of that paper was the non-dilatonic branes, for which
189: the energy density above extremality $\Delta m$ and entropy density
190: $s$ were written in a way reminiscent of gases of weakly interacting
191: massless particles in $d=p+1$.  For the dilatonic cases, however, the
192: expressions do not yield recognizable weak-coupling results. The
193: physical interpretation of this fact is that the strongly coupled
194: physics of the $p\not=3$ SYM theories gives rise to nontrivial
195: dependence of the entropy and the energy above extremality on the
196: temperature,
197: %
198: \begin{equation}\label{klebtseysimple}
199: \Delta m(T) \propto T^{\frac{2(7-p)}{(5-p)}} \,,\qquad s(T)
200: \propto T^{\frac{(9-p)}{(5-p)}} \,,
201: \end{equation}
202: %
203: where $T$ is the Hawking temperature of the geometry.
204: 
205: In the $p=0$ case, checking aspects of the open/closed-string
206: correspondence conjectures is potentially feasible.  The reason is of
207: course that the SYM theory in this case is actually matrix {\em
208: quantum mechanics} -- albeit with sixteen supercharges.  The numerical
209: investigations of \cite{kabat1,kabat2} aimed to check the $p=0$
210: correspondence explicitly, by finding the entropy of the system of
211: {\em strongly coupled} $N$ D0-branes and comparing it to the entropy
212: of the near extremal D0-brane supergravity background.  Results
213: obtained showed that the 9/5 power in the expression for the entropy
214: in (\ref{klebtseysimple}) was indeed approximately reproduced in the
215: numerical approach.  This striking result, in combination with the
216: work given a lightning review in the next subsection, provided the
217: essential motivation for our work.
218: 
219: %--------------------------------------------------------------------+
220: \subsection{Brane and antibranes at finite temperature and the entropy
221: of black branes}
222: %
223: Motivated by an observation of Horowitz, Maldacena, and
224: Strominger~\cite{Maldacena1}, Danielsson, Guijosa and Kruczenski
225: \cite{Danielsson} argued that, starting with a brane-antibrane system
226: at zero temperature, turning on a finite temperature could lead to
227: reappearance of open string modes.  This physics was argued to ensue
228: at strong open-string coupling $g_s N$ but weak closed-string coupling
229: $g_s$, as would be appropriate to a model with validity in the black
230: brane regime.  Note that, to have this reappearance of the open string
231: modes at temperatures {\em below} the Hagedorn temperature, going to
232: the regime of strong open-string coupling was crucial. In alternative
233: language, the tachyonic mode becomes stabilized by finite-temperature
234: and strong coupling physics. Based on these observations, the authors
235: of \cite{Danielsson} formulated a microscopic brane-antibrane model
236: for the entropy of non-dilatonic black branes: the D3, M2 and M5
237: cases.
238: 
239: The model has two SYM theories on worldvolumes, one for the branes and
240: the other for the anti-branes.  The tachyon, argued to be stabilized
241: by finite-temperature and strong coupling effects, does not contribute
242: to the total energy of the black hole, because it sits almost on top
243: of the tachyon potential. It is also argued that the tachyon does not
244: contribute to the total entropy, because the tachyon gets a large
245: thermal mass.
246: 
247: The black holes of interest to us are neutral.  Since the black
248: 3-brane in $D=10$ (or, equivalently, the black hole in $D=7$) is
249: neutral, it is taken to be modelled by the gas on a number $N$ of
250: D3-branes, the gas on an identical number $N$ ${\overline{\rm
251: D3}}$-branes, and the tachyon dynamics.  Since the branes and
252: anti-branes appear in equal numbers, it makes sense to assume that the
253: corresponding temperatures are equal.  Using these assumptions, for
254: the strongly coupled field theory side,
255: %
256: \begin{eqnarray}
257: m_{\rm FT} = 2N\tau_3+a\frac{\pi^2}{8}N^2 T^4, \nonumber\\ 
258: s_{\rm FT}=a\frac{\pi^2}{6}N^2T^3 \,,
259: \end{eqnarray}
260: %
261: where $\tau_3$ is the D3-brane tension and $a$ is a constant known
262: \cite{peet1} from AdS/CFT to be 6 (not 8) when the SYM theory in
263: question is strongly coupled.  Note that the quantities of
264: thermodynamic interest written here are specific, i.e.  mass and
265: entropy {\em densities}.  Rearranging to eliminate $T$ gives
266: %
267: \begin{equation}
268: s_{\rm FT} = a(\pi^2/6) N^2 \left(
269: {\frac{m_{\rm FT} - 2N\tau_3}{a(\pi^2/8)N^2}}
270: \right)^{3/4} \,.
271: \end{equation}
272: %
273: 
274: It is important to emphasize that this analysis is done in the {\em
275: microcanonical} ensemble, where the total mass and charge of the black
276: hole, as well as any angular momenta present, are kept fixed.  Working
277: in the microcanonical ensemble prevents thermal fluctuations from
278: creating an infinite number of D3-${\overline{\rm D3}}$ pairs.
279: Working in canonical picture, by contrast, would allow the system to
280: catastrophically create an infinite number of pairs, by extracting an
281: arbitrary amount of energy from the reservoir, as it is a
282: thermodynamically favourable state. In other words, one indefinitely
283: spends energy to make more entropy, which makes the thermal ensemble
284: destabilized.
285: 
286: The essential innovation in \cite{Danielsson} is that the total number
287: of branes is regarded as an independent variable, which is determined
288: thermodynamically.
289: %
290: The next step in the analysis, then, is to maximize the entropy with
291: respect to $N$.  The value of $N$ so obtained is related to the mass
292: density as
293: %
294: \begin{equation}
295: m_{\rm FT} = 5 N \tau_3 \,.
296: \end{equation}
297: %
298: 
299: To proceed further, it is necessary to decide what to do with the
300: masses.  One assumption is to take $m_{\rm FT}=m_{\rm SG}$.  Using
301: this, and substituting back for the optimized value of $N$ leads to
302: the final expression for the entropy from the strongly coupled field
303: theory side,
304: %
305: \begin{equation}
306: s_{\rm
307: FT}=a^{\frac{1}{4}}2^{\frac{5}{4}}3^{-\frac{1}{4}}5^{-\frac{5}{4}}
308: \pi^{\frac{1}{4}}\sqrt{\kappa} m_{\rm FT}^{\frac{5}{4}} \,,
309: \end{equation}
310: %
311: where $\kappa={\sqrt{\pi}}/{\tau_3}$. One the other hand, entropy of a
312: black 3-brane is
313: %
314: \begin{equation}
315: s_{\rm SG}=2^{\frac{9}{4}}5^{-\frac{5}{4}}\pi^{\frac{1}{4}}
316: \sqrt{\kappa}m_{\rm SG}^{\frac{5}{4}} \,.
317: \end{equation}
318: %
319: Identifying $m_{\rm SG}=m_{\rm FT}$ and $a=6$ gives
320: %
321: \begin{equation}
322: s_{\rm SG}= 2^{3/4} s_{\rm FT} \, .
323: \end{equation}
324: %
325: This was the result of \cite{Danielsson}.  It shows that the entropy
326: scaling from the brane-antibrane model is correct.  It is worth noting
327: here that this scaling agreement is nontrivial: it does {\em not} come
328: about through simple dimensional analysis (and counting powers of
329: $N$).
330: 
331: The overall coefficient misses, by a factor close to unity.  It is
332: perhaps not surprising, though, that the agreement, while close, is
333: not exact.  Reasons for this may include the fact that there is no
334: clean decoupling between open string and closed string degrees of
335: freedom.  These effects not taken into account in the brane-antibrane
336: model may indeed play a role, but they cannot be major effects for
337: quantities like the entropy, because the scaling comes out correct.
338: 
339: %====================================================================+
340: \section{D$p$ and ${\overline{\rm\bf D}p}$-branes
341: with angular momenta}
342: %
343: In order to test the idea of the brane-antibrane model for neutral
344: black hole entropy further, it is interesting to consider the cases
345: other than $p=3$, and to add angular momentum.  This will be the main
346: focus of the analytic part of our work.
347: 
348: It is convenient to begin by reviewing some salient properties of
349: rotating black D$p$-branes for various $p$.  We are particularly
350: interested in the $p=0$ story, since that is the case of the
351: brane-antibrane model which we plan to test numerically, starting with
352: the preliminary investigations of section 4.
353: 
354: %--------------------------------------------------------------------+
355: \subsection{Supergravity side}
356: 
357: Black $p$-branes in $D=10$ are of course equivalent to $D\equiv 10-p$
358: dimensional black holes.  We will not need the precise form of the
359: supergravity fields; what is important for us here is the relationship
360: between various physical parameters including tension, charge, angular
361: momenta, and horizon radius.  Using for example \cite{HarmarkObers},
362: and converting to quantities which are specific (per unit volume), we
363: have for the $D=10$ black branes
364: %
365: \begin{eqnarray}
366: %
367: j_i &=& {\frac{2}{(7-p)}}{\frac{\alpha_p}{16\pi G}}
368: r_0^{7-p}\ell_i {\frac{1}{\sqrt{1-\zeta^2}}} \,,\nonumber\\
369: %
370: T_H &=& {\frac{(7-p)-2\kappa}{4\pi r_H}} {\sqrt{1-\zeta^2}}
371: \quad {\rm{where}}\quad \kappa = \sum_{i=1}^n
372: {\frac{\ell_i^2}{\ell_i^2+r_H^2}} \,,\nonumber\\
373: %
374: m &=& {\frac{\alpha_p}{16\pi G}} r_0^{7-p} \left[
375: {\frac{1}{(7-p)}} + {\frac{1}{1-\zeta^2}} \right] \,,\nonumber\\
376: %
377: s &=& {\frac{4\pi}{(7-p)}} {\frac{\alpha_p}{16\pi G}} r_0^{7-p}
378: r_H {\frac{1}{\sqrt{1-\zeta^2}}} \,,\nonumber\\
379: %
380: q &=& {\frac{\alpha_p}{16\pi G}} r_0^{7-p} \left[
381: {\frac{\zeta}{1-\zeta^2}} \right] \,,
382: %
383: \end{eqnarray}
384: %
385: where $\zeta$ is the boost rapidity parameter while $r_0,\ell_i$ are
386: the original parameters used in creating the solutions, while the
387: Newton constant is given by $16\pi G = (2\pi)^{7} g_s^2 \ell_s^8$.  We
388: also use the shorthand $ \alpha_p \equiv (7-p)\Omega_{8-p}$.  Values
389: of $\alpha_p$ are tabulated here for convenience
390: %
391: \begin{center}
392: \begin{tabular}{|c|c|c|c|c|c|c|c|}\hline
393: $p$ & 0 & 1 & 2 & 3 & 4 & 5 & 6
394: %
395: \cr\hline
396: %
397: $\alpha_p$ & $7\pi^4/3$ & $32\pi^3/5$ & $5\pi^3$ & $32\pi^2/3$ &
398: $6\pi^2$ & $8\pi$ & $2\pi$
399: %
400: \cr\hline
401: \end{tabular}
402: \end{center}
403: %
404: The parameter $r_H$ is the horizon radius given by
405: %
406: \begin{equation}\label{r0rH}
407: r_H^{7-p} \prod_{i=1}^{n}\left( 1 + {\frac{\ell_i^2}{r_H^2}} \right) -
408: r_0^{7-p} = 0 \,.
409: \end{equation}
410: %
411: It is straightforward to massage the expressions to write the entropy
412: density $s$ in terms of the energy density above extremality $\Delta
413: m$.  To do that, it is convenient use (\ref{r0rH}) to obtain $r_0$ in
414: terms of $r_H$ and $\ell_i$.  This is simple if we make the
415: definitions
416: %
417: \begin{equation}
418: \rho \equiv {\frac{r_H}{r_0}} \ \,, \quad
419: \lambda_i \equiv {\frac{\ell_i}{r_H}} \,.
420: \end{equation}
421: %
422: Then we have
423: %
424: \begin{equation}\label{rhoeqn}
425: \rho(\lambda_i) \equiv {\frac{r_H}{r_0}} =
426: \left[ \prod_{i=1}^n
427: (1+\lambda_i^2) \right]^{-{\frac{1}{(7-p)}}} \,.
428: \end{equation}
429: %
430: The energy density above extremality $\Delta m \equiv m-q$ is given by
431: %
432: \begin{equation}\label{deltaE}
433: \Delta m = {\frac{\alpha_p}{16\pi G}} r_0^{7-p} \left[
434: {\frac{1}{(1+\zeta)}} + {\frac{1}{(7-p)}} \right] \,.
435: \end{equation}
436: %
437: We can also write the Hawking temperature as
438: %
439: \begin{equation}
440: T_H = \sqrt{1-\zeta^2} \, {\frac{(7-p)}{4\pi r_0}}
441: {\frac{k_p(\lambda_i)}{\rho(\lambda_i)}} \,,
442: \end{equation}
443: %
444: where $\rho(\lambda_i)$ is given by (\ref{rhoeqn}) and we use the
445: additional shorthand
446: %
447: \begin{equation}
448: k_p(\lambda_i) \equiv 1 - {\frac{2}{(7-p)}}\sum_{i=1}^n
449: {\frac{\lambda_i^2}{(1+\lambda_i^2) }} \,.
450: \end{equation}
451: %
452: In addition, the number of D$p$-branes can be written as
453: %
454: \begin{equation}\label{Npzeta}
455: N_p = {\frac{\alpha_p}{16\pi G}} {\frac{1}{\tau_p}} r_0^{7-p}
456: {\frac{\zeta}{1-\zeta^2}} \,,
457: \end{equation}
458: %
459: where
460: %
461: \begin{equation}
462: \tau_p = {\frac{1}{g_s(2\pi)^p\ell_s^{p+1}}}
463: \end{equation}
464: %
465: is the D$p$-brane tension.
466: 
467: Our main interest is to explain the entropy of neutral black branes in
468: $D=10$, or equivalently, neutral black holes in $D=10-p$.  We
469: therefore record the expressions here that we aim to reproduce using
470: the strongly coupled brane-antibrane field theory model.
471: 
472: In the case of a neutral black hole, we have $\Delta m = m_{\rm SG}$.
473: Defining
474: %
475: \begin{equation}
476: \delta_p \equiv {\frac{(8-p)}{(7-p)}} \,,
477: \end{equation}
478: %
479: we have for the energy density
480: %
481: \begin{equation}
482: m_{\rm SG} = {\frac{\alpha_p}{16\pi G}} r_0^{7-p} \delta_p
483: \end{equation}
484: %
485: and for the Hawking temperature
486: %
487: \begin{equation}
488: T_H = {\frac{(7-p)}{4\pi r_0}}
489: {\frac{k_p(\lambda_i)}{\rho(\lambda_i)}} \,.
490: \end{equation}
491: %
492: Therefore,
493: %
494: \begin{equation}
495: m_{\rm SG} (T_H) = {\frac{\alpha_p}{16\pi G}} \delta_p \left(
496: {\frac{(7-p)}{4\pi}} {\frac{k_p(\lambda_i)}{\rho(\lambda_i)}}
497: {\frac{1}{T_H}} \right)^{7-p} \,,
498: \end{equation}
499: %
500: while for the entropy
501: %
502: \begin{equation}
503: s_{\rm SG} (T_H) = {\frac{4\pi}{(7-p)}} {\frac{\alpha_p}{16\pi G}}
504: \rho(\lambda_i) \left( {\frac{(7-p)}{4\pi}}
505: {\frac{k_p(\lambda_i)}{\rho(\lambda_i)}} {\frac{1}{T_H}} \right)^{8-p}
506: \,.
507: \end{equation}
508: %
509: Defining
510: %
511: \begin{equation}
512: a_p \equiv {\frac{(7-p)}{2}} \delta_p^{\delta_p}
513: \left({\frac{\alpha_p}{16\pi}}\right)^{\delta_p-1} \,,
514: \end{equation} 
515: %
516: we have that
517: %
518: \begin{equation}
519: s_{\rm SG}(m_{\rm SG}) = {\frac{2\pi}{Ga_p}} \rho(\lambda_i)
520: (Gm_{\rm SG})^{\delta_p} \,.
521: \end{equation}
522: %
523: 
524: We may now ask how to unfurl the dependence of $\rho(\lambda_i)$ on
525: $m_{\rm SG}$ and the rotation parameters $j_i$.  The simplest way to
526: proceed is to recognize that
527: %
528: \begin{equation}
529: {\frac{\lambda_i}{2\pi}} = {\frac{j_i}{s}} \,.
530: \end{equation}
531: %
532: We therefore have the equation
533: %
534: \begin{eqnarray}\label{sugralambda}
535: \lambda_i \left[\prod_{i=1}^n (1+\lambda_i^2)
536: \right]^{-{\frac{1}{(7-p)}}} &=& a_p {\frac{j_i}{G^{\delta_p-1}m_{\rm
537: SG}^{\delta_p}}} \nonumber\\
538: %
539: &=& {\frac{j_i}{R_{\rm SG}(m_{\rm SG},G)}} \,,
540: \end{eqnarray}
541: %
542: where
543: %
544: \begin{equation}
545: R_{\rm SG} \equiv {\frac{1}{a_p}} G^{\delta_p-1} m_{\rm SG}^{\delta_p}
546: \,.
547: \end{equation}
548: %
549: Then, finally, we have
550: %
551: \begin{equation}\label{sSGfnofrho}
552: s_{\rm SG} = 2\pi \rho(R_{\rm SG}(m_{\rm SG},G),j_i) \, R_{\rm
553: SG}(m_{\rm SG},G) \,.
554: \end{equation}
555: %
556: 
557: We want to invert (\ref{sugralambda}) to unfurl the dependence of
558: $\lambda_i$ on $m_{\rm SG}, G, j_i$.  When no angular momenta are
559: present, we have of course that $\rho=1$.  In general, however, the
560: equation for $\lambda_i$ is a polynomial of degree $(7-p)$ in
561: $\lambda$, which is a quintic or worse for $p\leq 2$.  For now, we do
562: not concern ourselves with whether we can actually solve for $\rho$;
563: we just leave the dependence of $\rho$ on $(m_{\rm SG},G,j_i)$
564: implicit.  We also note that for $p=3$ and one angular momentum
565: parameter $j_1$, we can actually solve to find
566: %
567: \begin{equation}
568: \lambda_1 (p=3) =
569: \sqrt{2\sqrt{\chi}\left(\sqrt{\chi}+\sqrt{1+\chi}\right)}
570: \end{equation}
571: %
572: and
573: %
574: \begin{equation}
575: \rho(p=3) = {\frac{1}{\sqrt{\sqrt{\chi}+\sqrt{1+\chi}}}} \,,
576: \end{equation}
577: %
578: where
579: %
580: \begin{equation}
581: \chi = 2^{-5} 3^{-1} 5^5 \pi {\frac{j_1^4}{G\,m_{\rm SG}^5}} \,.
582: \end{equation}
583: %
584: 
585: %--------------------------------------------------------------------+
586: \subsection{Near-extremal: the field theory side}
587: %
588: In this subsection, we take the near-extremal limit of various
589: supergravity formul\ae\ to tell us how the strongly coupled field
590: theory quantities should behave on the branes and anti-branes, using
591: IMSY duality.
592: 
593: We should mention that in some respects our analysis here is quite
594: similar to \cite{KlebanovTseytlin}.  There are two major differences,
595: however. The first is that we are interested in the effect of turning
596: on angular momenta $j_i$.  Our results must of course reduce to those
597: of \cite{KlebanovTseytlin} upon taking $j_i=0$.  The second is that we
598: plan to use the information about the strongly coupled field theory,
599: obtained via weakly coupled supergravity, using the techniques of
600: \cite{Danielsson}.
601: 
602: Using the relation (\ref{Npzeta}), we can\footnote{The constant $c_p$
603: familiar from the TASI-99 lectures of one of us is related to the
604: quantities used here by $c_p g_s \ell_s^{7-p} = 16\pi G/\alpha_p$.}
605: express $r_0$ in terms of $N_p$
606: %
607: \begin{equation}\label{r0Npne}
608: r_0^{7-p} = 2(1-\zeta) \, N_p \tau_p {\frac{16\pi G}{\alpha_p}} \,,
609: \end{equation}
610: %
611: (Of course, this $r_0$ is appropriate to the near-extremal geometry,
612: and is not the same as the $r_0$ at the end of the previous
613: subsection, which refers to the horizon radius of the {\em neutral}
614: spacetime.) Equivalently,
615: %
616: \begin{equation}
617: 2(1-\zeta) = \left({\frac{r_0}{\ell_s}}\right)^{7-p}
618: {\frac{(2\pi\alpha_p)(2\pi)^{2(p-5)}}{\ell_s^{3-p}(g_{\rm YM}^2N)}}
619: \,.
620: \end{equation}
621: %
622: The energy density above extremality becomes
623: %
624: \begin{equation}
625: \Delta m_{\rm branes} = (1-\zeta) N_p \tau_p {\frac{(9-p)}{(7-p)}} \,.
626: \end{equation}
627: %
628: Near extremality we have\footnote{For reasons discussed in \cite{peet}
629: etc., the cases $p=5,6$ are more problematic to interpret.  We
630: therefore restrict ourselves from here on to the cases $p\leq 4$.} for
631: the Hawking temperature
632: %
633: \begin{eqnarray}\label{THzetane}
634: T_H &=& [2(1-\zeta)]^{\frac{(5-p)}{2(7-p)}} \left(
635: {\frac{\alpha_p}{16\pi GN_p\tau_p}} \right)^{\frac{1}{(7-p)}}
636: {\frac{(7-p)}{4\pi}} {\frac{k_p(\lambda_i)}{\rho(\lambda_i)}}
637: \nonumber\\
638: %
639: &=& [2(1-\zeta)(2\pi)^4]^{\frac{(5-p)}{2(7-p)}}
640: \left[{\frac{(2\pi\alpha_p)}{(g_{\rm
641: YM}^2N)\ell_s^4}}\right]^{\frac{1}{(7-p)}} {\frac{(7-p)}{4\pi}}
642: {\frac{k_p(\lambda_i)}{\rho(\lambda_i)}} \,.
643: \end{eqnarray}
644: %
645: We can obtain the equation of state for the near-extremal black
646: D$p$-brane by eliminating $\zeta$ in favour of $T_H$,
647: %
648: \begin{eqnarray}\label{ne_energy}
649: \Delta m_{\rm branes}(T_H) &=& N^2 \left\{ \gamma_p (2\pi)^2
650: \left(2\pi\alpha_p\right)^{-\frac{2}{(5-p)}} \right\} \times
651: \nonumber\\
652: %
653: && \times\quad \left(g_{\rm{YM}}^2N\right)^{\frac{(p-3)}{(5-p)}}
654: \left[ {\frac{4\pi}{(7-p)}} {\frac{\rho(\lambda_i)}{k_p(\lambda_i)}}
655: T_H \right]^{\frac{2(7-p)}{(5-p)}} \,.
656: \end{eqnarray}
657: %
658: Defining the abbreviation
659: %
660: \begin{equation}
661: \gamma_p \equiv {\frac{(9-p)}{2(7-p)}} \,,
662: \end{equation}
663: %
664: we have for the entropy density on the branes
665: %
666: \begin{eqnarray}\label{ne_entropy}
667: s_{\rm branes}(T_H) &=& N^2 \left\{ {\frac{4\pi}{(7-p)}} (2\pi)^{2}
668: \left(2\pi\alpha_p\right)^{-\frac{2}{(5-p)}} \right\} \times
669: \nonumber\\ 
670: %
671: && \times\quad \left( g_{\rm YM}^2N \right)^{\frac{(p-3)}{(5-p)}} \,
672: \rho(\lambda_i) \, \left[ {\frac{4\pi}{(7-p)}}
673: {\frac{\rho(\lambda_i)}{k_p(\lambda_i)}} T_H
674: \right]^{\frac{(9-p)}{(5-p)}} \,.
675: \end{eqnarray}
676: %
677: This equation is of central importance; it encodes the equation of
678: state for the system.
679: 
680: Now we come to the crucial step.  Following the innovation in
681: \cite{Danielsson}, we actually take the lead for the behaviour of the
682: strongly coupled field theory on the branes and anti-branes by using
683: the near-extremal supergravity results.  This is tantamount to using
684: IMSY duality \cite{itzhaki}.  Here, for general $p$, our proposal to
685: use IMSY duality for the $p\not=3$ case in the model of type
686: \cite{Danielsson} is a more nontrivial step than in the $p=3$ case
687: where the field theory behaved in a simple way.  This assumption
688: involves some nontrivial physics; like \cite{Danielsson}, we are not
689: taking into account the lack of a clean decoupling limit in the
690: brane-antibrane model.  Nonetheless, it is our point of view here that
691: taking the near-extremal supergravity result seriously for the
692: strongly coupled field theory on both the set of branes and the set of
693: antibranes is exactly what we need.
694: 
695: We are particularly interested to investigate this story for the
696: D0-${\overline{\rm{D}}0}$ case. The reason is that in the $d=0+1$ case
697: we have the hope of actually checking the above assumption explicitly,
698: using a strong-coupling simulation. In particular, a significant
699: motivation for our investigation of the non-conformal ($p\not=3$)
700: cases in the first place was the result of \cite{kabat1,kabat2} in
701: which the equation of state was approximately reproduced numerically.
702: We will begin developing the numerical story for the
703: D0-${\overline{\rm{D}}0}$ case in the next section, but for now we
704: work out the analytics.
705: 
706: For the open-string gas on (say) the set of D$p$-branes, we have
707: %
708: \begin{equation}
709: s_{\rm branes} = {\frac{2\pi}{b_p}} \rho(\lambda_i) \sqrt{N} g_{\rm
710: YM}^{\frac{(p-3)}{(7-p)}} \left(\Delta m\right)^{\gamma_p} \,,
711: \end{equation}
712: %
713: where we have defined
714: %
715: \begin{equation}
716: b_p \equiv {\frac{(7-p)}{2}} \gamma_p^{\gamma_p}
717: (2\pi)^{\frac{(p-4)}{(7-p)}}
718: \alpha_p^{\frac{1}{(7-p)}} \,.
719: \end{equation}
720: %
721: 
722: Now, for our model, recalling that we have the field theory on both
723: the branes {\em and} antibranes, we have
724: %
725: \begin{equation}\label{eTHne}
726: m_{\rm FT} = (2)N \tau_p + (2)\Delta m(T_{\rm FT}) \,.
727: \end{equation}
728: %
729: BPS branes carry no macroscopic entropy, so we take the entropy of the
730: strongly coupled field theory system representing the neutral black
731: brane to be twice the entropy of the gas on each set of branes,
732: $s_{\rm FT}(T_{\rm FT}) = (2)s(T_{\rm FT})$.  We find
733: %
734: \begin{equation}\label{SFTrot}
735: s_{\rm FT} = (2) {\frac{2\pi}{b_p}} \rho(\lambda_i) \sqrt{N} \,
736: g_{\rm{YM}}^{\frac{(p-3)}{(7-p)}} \, \left( {\frac{m_{\rm
737: FT}}{(2)}}-N\tau_p\right)^{\gamma_p} \,.
738: \end{equation}
739: %
740: We now need to know how $\rho(\lambda_i)$ depends on variables of
741: interest.  We again use the trick of the previous subsection to write
742: $\lambda_i = 2\pi j_i/s$.  Then
743: %
744: \begin{eqnarray}\label{lambdabp}
745: \lambda_i\rho(\lambda_i)= \lambda_i \left[\prod_{i=1}^n
746: (1+\lambda_i^2) \right]^{-{\frac{1}{(7-p)}}} &=& {\frac{b_p}{(2)}}
747: {\frac{1}{\sqrt{N}}} \left(g_{\rm YM}\right)^{\frac{(3-p)}{(7-p)}} j_i
748: \left( {\frac{m_{\rm{FT}}}{(2)}} -N\tau_p\right)^{-\gamma_p}
749: \nonumber\\ &=& {\frac{j_i}{R_{{\rm FT}}(N, m_{\rm FT}, g_{\rm
750: YM},\tau_p)}} \,,
751: \end{eqnarray}
752: %
753: where
754: %
755: \begin{equation}\label{RFTdefn}
756: R_{{\rm FT}}(N,m_{\rm FT},g_{\rm YM},\tau_p) = (2) {\frac{1}{b_p}}
757: \sqrt{N}g_{\rm YM}^{\frac{(p-3)}{(7-p)}} \left(
758: {\frac{m_{\rm{FT}}}{(2)}} -N\tau_p\right)^{\gamma_p} \,.
759: \end{equation}
760: %
761: Referring back to the entropy equation (\ref{SFTrot}), we have for
762: general $p$ and general angular momenta that
763: %
764: \begin{equation}\label{sFTfnofrho}
765: s_{\rm FT} = 2\pi \, R_{{\rm FT}}(N,m_{\rm FT},g_{\rm YM},\tau_p)
766: \, \rho(R_{{\rm FT}}(N,m_{\rm FT},g_{\rm YM},\tau_p),j_i) \,.
767: \end{equation}
768: %
769: 
770: At this stage, we may wonder whether the equations for the physical
771: parameters $\lambda_i$ in terms of $(j_i,m_{\rm FT},N,g_{\rm
772: YM},\tau_p)$ can actually be solved explicitly analytically.  For some
773: $p$, they can.  For other cases, including the $p=0$ case of most
774: interest to us, however, they cannot.  For now, we will put this issue
775: aside, and just proceed with the dependence of $\rho$ on $R_{\rm FT}$
776: and thereby on $(N,m_{\rm FT}, g_{\rm YM},\tau_p, j_i)$ implicit.
777: 
778: %--------------------------------------------------------------------+
779: \subsection{Comparing field theory and supergravity}
780: %
781: We now take the field theory result of the last subsection and ask
782: what happens when we optimize with respect to $N$, the number of
783: branes (and anti-branes).  For simplicity, we first ask how this story
784: works without angular momenta.  Later we add angular momenta back in.
785: 
786: Optimizing the entropy of the strongly coupled brane-antibrane
787: theories w.r.t. $N$ gives
788: %
789: \begin{equation}
790: N = {\frac{m_{\rm FT}}{2\tau_p(1+2\gamma_p)}} \, \quad {\rm i.e.}\quad
791: m_{\rm FT} = (2N\tau_p) 2\delta_p \,.
792: \end{equation}
793: %
794: The energy density in the brane gases is then
795: %
796: \begin{equation}\label{egasebranes}
797: m_{\rm FT} - 2N\tau_p = (2N\tau_p)2\gamma_p \,.
798: \end{equation}
799: %
800: Substituting these expressions for $N$ and $m_{\rm FT}$ back into the
801: expression for the entropy, and converting D$p$-brane quantities into
802: the Newton constant, we obtain
803: %
804: \begin{equation}
805: s_{\rm FT} = 2^{-\gamma_p} {\frac{2\pi}{a_p}}
806: G^{\delta_p-1}m_{\rm FT}^{\delta_p} \,.
807: \end{equation}
808: %
809: Two possible conclusions can be drawn from this.  The first is that
810: %
811: \begin{equation}
812: m_{\rm FT} = m_{\rm SG} \quad{\rm{and}}\quad s_{\rm
813: FT}=2^{-\frac{(9-p)}{2(7-p)}} \, s_{\rm SG}
814: \end{equation}
815: %
816: Alternatively, we can conclude that
817: %
818: \begin{equation}
819: s_{\rm FT} = s_{\rm SG} \quad{\rm{and}}\quad m_{\rm FT} =
820: 2^{\frac{(9-p)}{2(8-p)}} \, m_{\rm SG} \,.
821: \end{equation}
822: %
823: The second interpretation has an interesting conclusion. It says that
824: the field theory energy, which is simply the energy on the D$p$-branes
825: plus the energy on the ${\overline{{\rm D}p}}$-branes, is not simply
826: the supergravity energy, but there is a (suitably negative) binding
827: energy
828: %
829: \begin{equation}
830: \left| m_{\rm binding} \right| = m_{\rm SG} \left(
831: 2^{\frac{(9-p)}{2(8-p)}} -1 \right) \,.
832: \end{equation}
833: %
834: We find this interpretation the more attractive one.  Binding energy
835: can be expected in our model, because of the lack of a {\em clean}
836: decoupling limit in our system between the open-string and
837: closed-string modes.
838: 
839: Let us now add back the angular momenta and see if it affects our
840: exposition of the basic physics of the brane-antibrane model.
841: 
842: Recall that, in the microcanonical ensemble that we are studying, both
843: $m_{\rm FT}$ and $j_i$ are constants.  This is why we have labelled
844: $R_{\rm FT}(N$) a function of $N$, which we have to optimize following
845: the model of $\cite{Danielsson}$.
846: 
847: Regardless of whether the equation for $\lambda_i$ can be solved
848: explicitly, the entropy depends only on $\rho(R_{\rm FT})$, as in
849: (\ref{sFTfnofrho}). So we just proceed with $\rho(R_{\rm FT})$
850: implicit.
851: 
852: Our principle is to maximize the entropy as a function of $N$.  We
853: have
854: %
855: \begin{eqnarray}\label{maxiboo}
856: {\frac{1}{s_{\rm FT}}}{\frac{\partial s_{\rm FT}}{\partial N}} &=&
857: {\frac{1}{R_{\rm FT}}} {\frac{\partial R_{\rm FT}}{\partial N}} +
858: {\frac{1}{\rho(R_{\rm FT})}} {\frac{\partial\rho}{\partial R_{\rm
859: FT}}} {\frac{\partial R_{\rm FT}}{\partial N}} \nonumber\\
860: %
861: &=& \left[ {\frac{1}{R_{\rm FT}}} + {\frac{1}{\rho(R_{\rm FT})}}
862: {\frac{\partial\rho}{\partial R_{\rm FT}}} \right] {\frac{\partial
863: R_{\rm FT}}{\partial N}} \,.
864: \end{eqnarray}
865: %
866: Since we do not in general know the analytic dependence of $\rho$ on
867: $R_{\rm FT}$, we need to find the derivative implicitly.  We have
868: from (\ref{RFTdefn}) 
869: %
870: \begin{equation}
871: \lambda_i {\frac{1}{\rho}} {\frac{\partial\rho}{\partial R}} =
872: -{\frac{j_i}{R^2\rho}} - {\frac{\partial\lambda_i}{\partial R}} \,.
873: \end{equation}
874: %
875: Using (\ref{sFTfnofrho}), and $\lambda_i/(2\pi)=j_i/s$, we have
876: $\lambda_i=j_i/(\rho R)$, which gives
877: %
878: \begin{equation}
879: \lambda_i\left[ {\frac{1}{R_{\rm FT}}} + {\frac{1}{\rho(R_{\rm FT})}}
880: {\frac{\partial\rho}{\partial R_{\rm FT}}} \right] = -
881: {\frac{\partial\lambda_i}{\partial R_{\rm FT}}}
882: \end{equation}
883: %
884: Referring back to (\ref{lambdabp}), it is easy to find $\partial
885: \lambda_i/\partial R$,
886: %
887: \begin{equation}
888: {\frac{\left[(7-p)+(5-p)\lambda_i^2 \right]\lambda_i^{6-p}}{
889: (1+\lambda_i^2) \prod_j(1+\lambda_j^2) }} 
890: {\frac{\partial\lambda_i}{\partial R_{\rm SG}}} =
891: -{\frac{(7-p)}{R_{\rm SG}^{8-p}}} j_i^{7-p}
892: \end{equation}
893: %
894: Therefore the derivative is manifestly negative, as long as $p\leq 4$,
895: and it is nonzero as long as at least one angular momentum parameter
896: ($j_i$) is turned on.  Consequently, the term in square brackets in
897: (\ref{maxiboo}) does not vanish, and it is nonsingular.
898: %
899: Therefore, whatever the behaviour of $\rho(R_{\rm FT})$, we are safe
900: in concluding that the entropy is extremized by demanding that
901: $\partial R_{\rm FT}(N)/\partial N=0$.  Optimizing $R_{\rm FT}(N)$ we
902: find that
903: %
904: \begin{equation}
905: N = {\frac{m_{\rm FT}}{2\tau_p(1+2\gamma_p)}} \,.
906: \end{equation}
907: %
908: which is exactly what we had for the non-rotating case.
909: 
910: Substituting back this optimal value of $N$ into the field theory
911: quantity $R_{\rm FT}$, we find that
912: %
913: \begin{equation}
914: R_{\rm FT}(m_{\rm FT},j_{i\,{\rm FT}},g_{\rm YM},\tau_p)
915: = R_{\rm SG}(m,j_{i\,{\rm SG}},G) \,,
916: \end{equation}
917: %
918: where $j_{i\,{\rm FT}}$ are the angular momenta in {\em one} copy of
919: the strongly coupled field theory, if we make the identifications
920: %
921: \begin{equation}\label{idees}
922: m_{\rm FT} = m_{\rm SG} 2^{\frac{(9-p)}{2(8-p)}} \,,\quad
923: j_{i\,{\rm SG}} = j_{i\,{\rm FT}} \,.
924: \end{equation}
925: %
926: Now, na\"ively we would have expected the total angular momenta of the
927: neutral supergravity solution to be split in half, shared equally
928: between the strongly coupled brane and antibrane field theories.  The
929: fact that the angular momenta and the mass do not match precisely is
930: not particularly surprising, however, because of the lack of a clean
931: decoupling limit.  In fact, closed-string modes (whose physics not
932: included in the brane-antibrane model) might carry both mass and
933: angular momenta.  We find it intriguing, though, that the
934: renormalization factors are simply powers of two!
935: 
936: Now, in order to get the entropy to match between the field theory
937: side and the supergravity side here, we have the freedom to apply
938: renormalization factors to both the mass and angular momenta.
939: Alternatively, there is insufficient information to set these
940: renormalizations factors unambiguously using our analysis. 
941: 
942: Let us then accept the mass and angular momenta renormalizations of
943: (\ref{idees}).  We then find the remarkable fact that the functional
944: form of the field theory entropy (\ref{sFTfnofrho}) in terms of $R$ is
945: identical to the functional form of the entropy on the supergravity
946: side (\ref{sSGfnofrho}).  With the renormalizations (\ref{idees}), we
947: see that the $R$'s are the same on both sides.  The strongly coupled
948: field theory of branes and antibranes therefore reproduces the
949: supergravity result, up to renormalization factors of two, i.e.
950: %
951: \begin{equation}
952: s_{\rm FT}\left(m_{\rm FT}=2^{{\frac{(9-p)}{2(8-p)}}} m_{\rm
953: SG},j_{i\,FT}  \right) = s_{\rm SG}(m_{\rm SG},j_{i\,SG}) \,.
954: \end{equation}
955: %
956: This is our main analytic result. It shows that, regardless of
957: rotation or $p$, the entropy of the neutral black brane can be
958: recovered from the strongly coupled brane-antibrane field theory.  It
959: is also worth noting that this result {\em cannot} be obtained just by
960: dimensional analysis.  
961: 
962: Alternatively, we can think of this result as constituting a highly
963: nontrivial relationship between the thermodynamical properties of
964: near-extremal black branes and those of neutral black branes.
965: 
966: Before moving on to develop some more physics of the brane-antibrane
967: model, we may ask what our conclusion about the mass renormalization
968: may do to other parts of our brane-antibrane-gas model.  We know that
969: %
970: \begin{equation}
971: T_H^{-1} = {\frac{\partial s_{\rm SG}}{\partial m_{\rm SG}}} \,.
972: \end{equation}
973: %
974: Therefore if, as we have shown, the entropies match but the masses are
975: not equal, then the temperature is also affected in the same
976: proportion as the mass
977: %
978: \begin{equation}
979: T_{\rm FT} = T_H 2^{\frac{(9-p)}{2(8-p)}} \,.
980: \end{equation}
981: %
982: (Note that this does not spoil any of our previous assumptions.)  
983: 
984: One might wonder how the unusual thermal properties of neutral black
985: holes and black branes, such as negative heat capacity, could possibly
986: be understood from analyzing this D$p$-${\overline{{\rm D}p}}$ field
987: theory model, which is based on ordinary super-Yang-Mills systems in
988: various dimensions. The explanation of this point in the context of
989: the D$p$-${\overline{{\rm D}p}}$ model is simple and was given for the
990: conformal cases in \cite{Danielsson}.  Upon inspection, one finds that
991: there is a correlation between the energy in the open string gas and
992: the contribution to the energy coming from the tension of the branes.
993: Actually they are proportional to each other. This would give rise to
994: the following interpretation for the negative specific heat.  Namely,
995: that since the energy in the gases is proportional to the tension
996: energy, if we add more energy to the open string gases then we must
997: create more D$p$-${\overline{{\rm D}p}}$ pairs to maintain the
998: proportionality.  This requirement to make more D$p$-${\overline{{\rm
999: D}p}}$ pairs makes the system cool down, meaning that the more massive
1000: the black branes the colder they get.  As we see, the moral of the
1001: story is that `normal' SYM field theories on the worldvolume of the
1002: branes and antibranes behave `abnormally', because the total number of
1003: degrees of freedom controlled by $N$ is not a constant; rather, it is
1004: given thermodynamically by the entropy maximization scheme.
1005: 
1006: So it is natural and important to ask whether the same kind of
1007: correlations also hold in our particular system of interest, i.e.
1008: D0-${\overline{{\rm D}0}}$, and in an even broader sense, for a
1009: general D$p$-${\overline{{\rm D}p}}$ system.
1010: 
1011: It is easy to show the linear proportionality (\ref{egasebranes})
1012: between the energy in the gas and the energy contribution from the
1013: brane tension.
1014: Using 
1015: data from the optimization of $N$
1016: %
1017: \begin{equation}
1018: N\sim {\frac{m_{\rm FT}}{\tau_p}} \,,
1019: \end{equation}
1020: %
1021: where $m_{\rm FT}= 2N\tau_p + m_{\rm gas}$, gives
1022: %
1023: \begin{equation}
1024: m_{\rm gas}\sim N\tau_p \,.
1025: \end{equation}
1026: %
1027: It is intriguing that exactly the same behaviour was observed in the
1028: case of non-dilatonic branes (D3, M2, M5).  However, the way the
1029: result of \cite{Danielsson} was obtained, comparing energy density and
1030: pressure in field theory and supergravity, does not apply to our $p=0$
1031: case of particular interest as there is no pressure on a $0+1$
1032: dimensional worldvolume.
1033: 
1034: It is important to note that in our entire analysis, the black objects
1035: under study are neutral - nonperturbatively nonextremal.  It is quite
1036: pleasing that the brane-antibrane model precisely explains the entropy
1037: of the neutral black branes and black holes, up to renormalizations of
1038: the mass and angular momenta that we computed.
1039: 
1040: %--------------------------------------------------------------------+
1041: \subsection{Horizon size}
1042: 
1043: It is interesting to ask whether the transverse fluctuations of the
1044: D$p$-$\overline{\rm D}p$ system in the microscopic picture can
1045: reproduce the size of the horizon of the corresponding black brane
1046: geometry.  We now do a scaling analysis, not keeping precise numerical
1047: factors.
1048: 
1049: Before addressing the general-$p$ cases, it is instructive to review
1050: the simplest case $p=3$.  For $N$ {\em near-extremal} D3-branes at
1051: strong coupling, power counting in $N$ and conformal symmetry tell us
1052: that\footnote{The r.m.s. expectation value is normalized with a factor
1053: of $1/N$.}  $\langle {\vec{X}}^2 \rangle_{\rm rms}\sim N T^2$.  Using
1054: the model of \cite{Danielsson} to find the optimal value of $N$ gives
1055: $N\sim m_{\rm FT}/\tau_3\sim m_{\rm gas}/\tau_3$.  Here we have used
1056: the information from the brane-antibrane model both to set the optimal
1057: value of $N$ and to learn that there is roughly the same amount of
1058: energy density in the branes and in the open-string gas on those
1059: branes.  Next, we use the assumption $m_{\rm SG}\sim m_{\rm FT}$, and
1060: the supergravity relationship between the Hawking temperature and the
1061: energy, $m_{\rm SG}\sim r_0^4/G$.  We also recall the assumption that
1062: the brane gas temperature is equal to the antibrane gas temperature,
1063: and both are equal to the Hawking temperature of the black brane.
1064: Collecting these facts, and using the relationship $G\tau_3^2\sim 1$,
1065: then gives a relationship between the supergravity horizon radius and
1066: variables in the field theory.  The last assumption to be used is the
1067: supergravity relationship between the Hawking temperature and horizon
1068: radius $T_H\sim 1/r_0$.  So $\langle {\vec{X}}^2 \rangle_{\rm rms}
1069: \sim r_0^4 (1/r_0)^2 \sim r_0^2$.  Therefore, the r.m.s. extent of the
1070: position fields of the field theory corresponds to the horizon radius.
1071: In other words, we started from a near-extremal gauge theory and ended
1072: up with the size of the horizon of a {\em neutral} black 3-brane.
1073: 
1074: Now we turn away from the conformal case.  In later sections, we will
1075: be particularly interested in the $p=0$ case, so let us look at it
1076: here in some detail.
1077: 
1078: In the mean field approximation, it has been shown \cite{KabatLowe}
1079: that the extent of the ground state of the D0-brane theory at low
1080: temperatures is controlled by t'Hooft coupling, {\em not} the
1081: temperature.  In supergravity, on the other hand, the rough estimate
1082: of the size {\em is} temperature dependent.  Therefore, the above
1083: logic that we used for the D3-brane case will not work for the
1084: D0-brane case. In \cite{KabatLowe}, a remedy for this confusing
1085: situation was proposed. Namely, that fast-varying degrees of freedom
1086: are physically inaccessible to a local supergravity observer. The
1087: supergravity probes just cannot resolve high-frequency fluctuations of
1088: order of t'Hooft energies, that are occurring at a ``microscopic''
1089: level in quantum mechanics.  In other words, a supergravity probe
1090: cannot resolve distances sharper than $l_{\rm probe}\sim \beta$. As a
1091: consequence, the supergravity probes simply miss all the quantum
1092: dynamics of the D0-brane dynamics involving energies much higher than
1093: the Hawking temperature.  Therefore, in making a microscopic model to
1094: reproduce supergravity, caution must be used regarding which degrees
1095: of freedom have to be included in the picture.
1096: 
1097: Practically, the suggestion of \cite{KabatLowe} amounts to imposing a
1098: temperature-dependent cutoff on the spectral density of the transverse
1099: fluctuations, i.e. $X^i$ propagators, that leads to a temperature
1100: dependent $\langle{\vec{X}}^2\rangle_{\rm rms}$ which qualitatively
1101: shows the same behavior as $r^2_0$ does on supergravity side.  
1102: 
1103: So in what follows we will assume that, if one tries hard from the
1104: gauge theory point of view, the extension of the wavefunction of the
1105: near-extremal geometry should become consistent with supergravity
1106: expectations.  We use this intuition motivated by IMSY in what
1107: follows.
1108: 
1109: We now apply this thinking that came from studying D0-brane physics to
1110: the general-$p$ case.  We put together a number of ingredients that we
1111: have discussed in this section.  We begin with
1112: %
1113: \begin{equation}
1114: m_{\rm FT}\sim m_{\rm branes} + m_{\rm
1115: gas} \,,
1116: \end{equation}
1117: %
1118: and note from the model of \cite{Danielsson} we have 
1119: %
1120: \begin{equation}
1121: m_{\rm gas}\sim m_{\rm branes} \,.
1122: \end{equation}
1123: %
1124: (Of course, the same energy density is in the brane gas and antibrane
1125: gas, so for the purposes of scaling we do not need to compute each
1126: contribution separately.)  
1127: 
1128: From e.g. \cite{itzhaki}, we have for the spatial extent of a {\em
1129: near-extremal} D$p$ or ${\overline{{\rm D}p}}$ geometry,
1130: %
1131: \begin{equation}
1132: U_0\sim (g_{\rm YM}^2 N)^{1/(5-p)} T^{2/(5-p)} \,,
1133: \end{equation}
1134: %
1135: where $U_0\equiv r_0/\ell_s^2$.  
1136: %
1137: (Thinking of $U_0$ as distance and $T$ as energy, we see that this is
1138: exactly the same energy/distance relation found by \cite{peet} for a
1139: $D=10$ supergraviton probe of a D$p$-brane geometry in the decoupling
1140: limit.  The coincidence is not too surprising.)
1141: %
1142: We now add in dynamical information from the brane-antibrane model,
1143: %
1144: \begin{equation}
1145: N\sim m_{\rm FT}/\tau_p \,,
1146: \end{equation}
1147: %
1148: to give
1149: %
1150: \begin{equation}
1151: r_0^{5-p}\sim G m_{\rm FT} T^{2/(5-p)} \,.
1152: \end{equation}
1153: %
1154: Lastly, we bring in the (field theory) relationship between
1155: temperature and mass
1156: %
1157: \begin{equation}
1158: T \sim (Gm_{\rm FT})^{-1/(7-p)} \,.
1159: \end{equation}
1160: %
1161: Finally, we find 
1162: %
1163: \begin{equation}
1164: r_0\sim (Gm_{\rm FT})^{1/(7-p)} \,.
1165: \end{equation}
1166: %
1167: Since $m_{\rm FT}\sim m_{\rm SG}$, this is the same radius as that of
1168: the horizon in the {\em neutral} geometry of interest.  Therefore, the
1169: brane-antibrane model does successfully provide a consistent picture
1170: of the spatial extent of the horizon.
1171: 
1172: We now turn to checking the consistency of the supergravity
1173: approximation itself.
1174: 
1175: %--------------------------------------------------------------------+
1176: \subsection{Validity of supergravity}
1177: 
1178: The black D$p$-branes on which we concentrate here are neutral.  In
1179: order that they can be thought of as bona fide supergravity entities,
1180: we require at a minimum that the curvature at the horizon radius
1181: should be small in string units, to keep $\alpha^\prime$ corrections
1182: small.  In scaling,
1183: %
1184: \begin{equation}
1185: (Gm)^{1/(7-p)} \gg \ell_s \,.
1186: \end{equation}
1187: %
1188: Using the equilibrium value of $N$, this becomes
1189: %
1190: \begin{equation}
1191: (g_{s}^2\ell_s^8N\tau_p)^{1/(7-p)}\gg \ell_s \,,
1192: \end{equation}
1193: %
1194: which in turn leads to
1195: %
1196: \begin{equation}
1197: g_{s}N \gg 1 \,.
1198: \end{equation}
1199: %
1200: Of course, we also require that string loop corrections be under
1201: control
1202: %
1203: \begin{equation}
1204: g_s \ll 1 \,,
1205: \end{equation}
1206: %
1207: for the $D=10$ neutral black brane spacetime.
1208: 
1209: Now, let us recall one important piece of physics from
1210: \cite{Danielsson}, the study of the conformal case.  Even though the
1211: neutral geometry of interest is not studied in the decoupling limit,
1212: the brane and antibrane systems in the microscopic model are actually
1213: taken to be decoupled -- as facts from AdS/CFT are used to describe
1214: the strongly coupled D3-brane theory.  Therefore, at this point, it is
1215: appropriate to check whether the conditions for validity of the $D=10$
1216: near-extremal D$p$-brane geometry are satisfied here also for the
1217: non-conformal cases, as these geometries dictate for us the behaviour
1218: of the strongly coupled field theories on the D$p$ and
1219: ${\overline{{\rm D}p}}$.
1220: 
1221: Let us therefore study the IMSY conditions carefully, to gain
1222: understanding of the brane-antibrane side of the picture.  As
1223: described in \cite{itzhaki}, there is a region in which type II $D=10$
1224: supergravity is valid
1225: %
1226: \begin{equation}\label{imsycondition}
1227: 1\ll g_{\rm eff}^2(U) \ll N^{\frac{4}{(7-p)}} \,,
1228: \end{equation}
1229: %
1230: where the effective coupling at energy scale $U$ is
1231: %
1232: \begin{equation}\label{boo1}
1233: g_{\rm eff}^2(U) \sim g_{\rm YM}^2 N U^{p-3} \,.
1234: \end{equation}
1235: %
1236: Also, the horizon radius for the near-extremal geometry is related to
1237: the energy above extremality $m_{\rm gas}$ and the temperature $T$ by
1238: %
1239: \begin{equation}\label{boo2}
1240: U_0 \sim (g_{\rm YM}^4 m_{\rm gas})^{\frac{1}{(7-p)}} \sim (g_{\rm
1241: YM}^2N)^{\frac{1}{(5-p)}} T^{\frac{2}{(5-p)}} \,.
1242: \end{equation}
1243: %
1244: 
1245: Now, our big neutral black brane will have a low Hawking temperature.
1246: This means that we are not in danger of violating the $\alpha^\prime$
1247: (left-hand) end of the bound (\ref{imsycondition}).  However,
1248: precisely because we are operating at such low temperatures, we may be
1249: concerned about violating the strong-coupling (right-hand) end of the
1250: bound (\ref{imsycondition}).  For example, for $p=0$, we may be
1251: concerned about having to lift up to $D=11$.
1252: 
1253: It is a satisfying fact that our microscopic brane-antibrane model
1254: remains consistent in the $D=10$ picture.  The essential reason for
1255: this is that, in the brane-antibrane model, $N$ is not an independent
1256: variable.  Rather, $N$ is thermodynamically determined.  Using the
1257: fact that $m_{\rm gas}\sim N\tau_p$ and formul\ae\
1258: (\ref{boo1},\ref{boo2}), and requiring that both ends of the IMSY
1259: bound (\ref{imsycondition}) are respected gives two conditions,
1260: which reduce to
1261: %
1262: \begin{equation}
1263: g_sN\gg 1 \,, \qquad g_s \ll 1 \,.
1264: \end{equation}
1265: %
1266: As we saw before, open strings are strongly coupled, but closed
1267: strings are weakly coupled.
1268: 
1269: Therefore, we do not need to be concerned about departing from the
1270: regime of validity of $D=10$ supergravity, for our systems of strongly
1271: coupled branes and antibranes.
1272: 
1273: In particular, for the $p=0$ case, we are always in the $D=10$
1274: supergravity regime.  This means, in particular, that we never get to
1275: the $D=11$ supergravity regime.  Therefore, it is not apparent whether
1276: there is any simple relationship between these microscopic models for
1277: $D=10$ neutral black branes and the BFSS matrix theory models, whose
1278: microphysical description is in terms of D0-brane degrees of freedom
1279: representing $D=11$ M theory.
1280: 
1281: %====================================================================+
1282: \section{Numerically investigating 
1283: ${\rm\bf D0}{-}\overline{\rm\bf D0}$ physics}
1284: 
1285: For the remainder of this paper we will concentrate on the $p=0$ case,
1286: i.e. the brane-antibrane model of the $D=10$ Schwarzschild black
1287: hole. The reason is that the physics is a quantum mechanics, lending
1288: itself to the possibility of actually computing the behaviour of the
1289: strongly coupled field theory.
1290: 
1291: %--------------------------------------------------------------------+
1292: \subsection{Scalings, and strategy}
1293: 
1294: Of course, the idea of doing direct numerical simulations in the
1295: strongly coupled QM is not new.  Kabat et al.~\cite{kabat1} have used
1296: a method called Variational Perturbation Theory (VPT) to check the
1297: IMSY gravity/gauge duality conjectures \cite{itzhaki} for the
1298: decoupling limit of D$p$-brane systems with sixteen supercharges.  In
1299: particular, the entropy of $N$ near extremal D0-branes was computed
1300: using VPT, and was found to match the with supergravity result to the
1301: level of approximation used \cite{kabat2}. The agreement is
1302: impressive.  In particular, the highly nontrivial temperature
1303: dependence of the free energy was obtained (approximately): $\beta
1304: F\propto T^{1.8}$!  Indeed, this could be considered as the first
1305: (approximate) nonperturbative check of any gauge theory/gravity
1306: duality.  The causal structure of spacetime from the point of view of
1307: the gauge theory also has been studied in the context of mean field
1308: Gaussian approximations~\cite{kabat3}.
1309: 
1310: The next step for our program to understand $D=10$ Schwarzschild black
1311: holes would be to justify the crude brane-antibrane picture drawn in
1312: the last few sections, by computing the microscopic entropy using the
1313: effective quantum description of a system of D0-$\overline{\rm D0}$s.
1314: 
1315: Now, in the D0-$\overline{\rm D0}$ system we cannot take a clean
1316: decoupling limit, if we expect to keep the open string tachyon and
1317: massless string modes but not massive string modes.  Therefore, our
1318: further progress in developing the brane-antibrane model is to be
1319: thought of as an approximate description, where the massive string
1320: modes are not fully decoupled.  Of course, the same story was true of
1321: the Danielsson et al. work \cite{Danielsson}.
1322: 
1323: Clearly, there are two different energy scales set by two different
1324: dimensionful coupling constants in our problem: $\alpha'$ and $g_{\rm
1325: YM}^2=g_{s}(2\pi)^{-2}\ell_s^{-3}N$. In gauge theory, everything is
1326: governed by $g_{\rm YM}^2$, while $\alpha'$ controls the mass of the
1327: tachyon.  For the physics of D0-branes alone, dimensional analysis and
1328: power counting in $N$ tells us \cite{kabat1} that a dimensionless
1329: quantity like $\beta F$ can be written as
1330: %
1331: \begin{equation}
1332: \beta F_{\rm D0} \propto N^2 \mathcal{F} (\frac{T}{(g_{\rm
1333: YM}^2N)^{1/3}}) \,.
1334: \end{equation}
1335: %
1336: Here, with the tachyon, we have two different scales: the t'Hooft
1337: coupling and $\alpha'$.  Thus, the free energy can be written as
1338: %
1339: \begin{eqnarray}
1340: \beta F_{{\rm D0}{-}\overline{\rm D0}} \propto N^2 \mathcal{G}
1341: (\frac{T}{(g_{\rm YM}^2N)^{1/3}}, \alpha'(g_{\rm YM}^2N)^{2/3})= N^2
1342: \mathcal{G} (\frac{T}{(g_{\rm YM}^2N)^{1/3}}, (g_{s}N)^{2/3}) \,.
1343: \end{eqnarray}
1344: %
1345: Therefore, the free energy written in dimensionless units will be a
1346: function of $g_{s}N$.  Of course, it will also depend on the
1347: dimensionless inverse temperature measured in 't Hooft units,
1348: $\tilde\beta$.
1349: 
1350: In the limit where massive open string excitations can be neglected,
1351: and the string coupling is small, the system is relatively simple.  We
1352: have two copies of the $D0$-brane theory - one for D0-branes and one
1353: $\overline{\rm D0}$-branes - plus a complex tachyon $T$ and a massless
1354: Majorana fermion $\Psi$ coming from D-$\overline{\rm D}$ open
1355: strings.
1356: 
1357: The strategy is to compute the free energy of this system as a
1358: function of the tachyon classical background field, plus other
1359: quantities at finite inverse temperature $\beta$.  In the field theory
1360: we compute at strong couplings using VPT, and in the closed string
1361: description the $D=10$ supergravity approximation is valid.  Then the
1362: following pieces of information could be read off immediately from the
1363: free energy
1364: %
1365: \begin{description}
1366: %
1367: \item[Tachyon static mass] The thermodynamically favourable value for
1368: the tachyon expectation value can be computed by minimizing the
1369: effective action, i.e. the free energy of the system as a function of
1370: the tachyon background.
1371: %
1372: \item[Sign of the Tachyon dynamical mass] Another important piece of
1373: information encoded in the free energy is the effective mass of the
1374: tachyon which is given by
1375: %
1376: \begin{equation}
1377: m^2_{T}=\beta\frac{\delta^2F}{\delta \langle T\rangle\delta \langle
1378: T\rangle^{\star}} \,,
1379: \end{equation}
1380: %
1381: where $\langle T\rangle$ is the tachyon expectation value.  This
1382: effective mass includes the contributions from infinite numbers of
1383: loop diagrams via the Schwinger-Dyson equation.
1384: %
1385: \item[Magnitude of the Tachyon Dynamical Mass] This quantity is a
1386: measure of the smallness of the tachyon fluctuations.  A large
1387: dynamical mass results in small contributions to the entropy from the
1388: tachyon.
1389: %
1390: \item[Phase Portrait of the theory] The final goal is to calculate the
1391: phase portrait of the system, as a function of inverse temperature
1392: $\beta$ and $g_sN$. A sign change in $m^2_{T}$ from negative to
1393: positive would be of great interest, since it would signal the tachyon
1394: stabilization phenomenon.  This stabilization would give a
1395: justification for why a gas of D0 and $\overline{\rm D0}$ branes does
1396: not simply annihilate all the way down to a bunch of closed strings.
1397: \end{description}
1398: 
1399: One note of caution.  The Hagedorn phenomenon in string theory might
1400: impact us here in one place: at temperatures of order of the string
1401: scale, massive open string excitations become important, but we are
1402: not incorporating massive excitations into our dynamics. In order to
1403: have a self-consistent description of the phenomena, therefore, we
1404: will look for a possible finite-temperature tachyon stabilization at a
1405: temperature well below the Hagedorn temperature.
1406: 
1407: We now turn to describing the technology which we will use to perform
1408: the numerical simulation of the D0-${\overline{{\rm D}0}}$ system.
1409: 
1410: %--------------------------------------------------------------------+
1411: \subsection{Variational perturbation theory (VPT)}
1412: %
1413: The basic idea of VPT is simple.  Using VPT to simulate
1414: strong-coupling D0-brane physics was first introduced in
1415: \cite{kabat1,kabat2}, and we review the salient points here for
1416: refernence.  Suppose that the theory of interest has action $S$, and
1417: the aim is to approximate its free energy at {\em strong couplings}.
1418: The idea is to use a free theory, with action $S_0$, with arbitrary
1419: tunable parameters.  The next step is to find values for these
1420: parameters in such a way the free theory is a best fit for the full
1421: interacting theory.
1422: 
1423: For any arbitrary $S_{0}$ the following identity holds
1424: %
1425: \begin{equation}
1426: \beta F=\beta F_{0} -\langle e^{-(S-S_{0})}-1\rangle_{0,C} \,,
1427: \end{equation}
1428: %
1429: where $0$ refers to the free theory and $C$ stands for connected
1430: contributions. Expanding the identity,
1431: %
1432: \begin{equation}
1433: \beta F=\beta F_{0}+\langle S-S_{0}\rangle_{0}-\frac{1}{2}
1434: \langle(S-S_{0})^2\rangle_{0,C}+ \cdots \,.
1435: \end{equation}
1436: %
1437: It is important to note that this expansion is {\em not} a
1438: perturbative expansion in the couplings of the original interacting
1439: theory of interest.
1440: 
1441: Now, if terms to all orders in the above expansion are kept, then
1442: there is of course no dependence on the parameters of the trial free
1443: theory, as it is an identity true for any $S_{0}$.  Taking a practical
1444: approach and terminating the series at any finite order, however, the
1445: series depends on the variational parameters.
1446: 
1447: The next step is to fix the variational parameters.  Minimizing the
1448: free energy with respect to those parameters (which is equivalent to
1449: requiring the trial free action to satisfy the Schwinger-Dyson
1450: equation) leads to a set of algebraic coupled equations (in general
1451: infinite in number), called ``gap equations".  Solution of the gap
1452: equations yields the variational parameters, which are then
1453: substituted back to obtain the free energy.
1454: 
1455: This VPT method has in fact been checked explicitly for quantum
1456: mechanical systems where calculations in the full interacting theory
1457: can actually be done exactly, and the above expansion captures the
1458: strong-coupling behaviour accurately; in particular, convergence is
1459: very fast.  Our system of interest here is of course not solvable, so
1460: we need to use the approximate expansion procedure outlined above.
1461: 
1462: Practically speaking, of course, the infinite set of equations cannot
1463: be solved.  Therefore it is necessary to cut off that infinite set of
1464: equations, and find the common roots of the finite coupled set of
1465: algebraic equations.  Solving this system of equations is, at any
1466: rate, computationally very expensive.
1467: 
1468: The last step is then to substitute the solution of the gap equations
1469: back into the free energy, to get the sought-after dependence on the
1470: parameters of interest. For us, these parameters will be the
1471: dimensionless inverse temperature $\tilde\beta$, and the open-string
1472: coupling $g_sN$.
1473: 
1474: As outlined in \cite{kabat1,kabat2}, the VPT method is not
1475: straightforward when dealing with supersymmetry and gauge theory.
1476: Regarding gauge theory.  Proposing a Gaussian theory for a
1477: non-dynamical gauge field in 0+1 dimensions is delicate, as one cannot
1478: simply gauge it away at finite temperature. Gauging away the gauge
1479: field leaves an observable (a Wilson loop), made from the zero mode of
1480: the gauge field encircling the Euclidean time direction, as a remnant.
1481: The Gaussian theory to be used \cite{kabat1} is the one-plaquette
1482: model studied by Gross and Witten~\cite{Gross}.  Taylor-Slavnov
1483: identities, which are consequences of gauge symmetry and relate
1484: various correlation functions, get invalidated by the VPT expansion.
1485: Regarding supersymmetry.  It turns out to be impossible in general to
1486: come up with a Gaussian theory with general variational parameters
1487: respecting supersymmetry, without inclusion of a trial action for
1488: auxiliary fields.  A way to get around the problem is to work with the
1489: off-shell superspace formulation.
1490: 
1491: On the other hand, the VPT method has some special positive features
1492: as well.  One is that VPT automatically respects 't Hooft counting.
1493: Another is that VPT automatically cures infrared problems arising from
1494: having an infinite moduli space of vacua where D0-branes are far
1495: apart.  The contribution coming from the zero mode sector goes like
1496: $\mathcal{O}(N)$ and is therefore distinguishable from
1497: ${\mathcal{O}}(N^2)$ contributions.  
1498: 
1499: %--------------------------------------------------------------------+
1500: \subsection{Motivating the action}
1501: %
1502: According to \cite{Kraus} (see also \cite{Tadashi} for a similar
1503: results), the action for a system of D9-$\overline{\rm D9}$ with the
1504: $U(1)$ gauge fields living on its world-volume is given by the
1505: following\footnote{We are using the mostly plus convention for the
1506: metric signature.}
1507: %
1508: \begin{eqnarray} S&=&-2\tau_{9}\int d^{10}x \,
1509: \exp(-2\pi\alpha^{\prime}T^{\dagger}T)\biggl[ 1 + 8\pi\alpha^{\prime2
1510: }\ln(2)D_{\mu}T^{\dagger}D^{\mu}T + \nonumber\\
1511: %
1512: && \qquad\qquad\qquad\qquad\frac{(2\pi\alpha^{\prime})^{2}}{8}
1513: {\overline{F}}_{\mu\nu} {\overline{F}}^{\mu\nu}
1514: +\frac{(2\pi\alpha^{\prime})^{2}}{8} F_{\mu\nu}F^{\mu\nu}
1515: \ldots\biggr] \,,
1516: \end{eqnarray}
1517: %
1518: where we have used `bars' to denote quantities in the ${\overline{\rm
1519: D0}}$ sector and `no bars' to denote quantities in the D0 sector.  Of
1520: course, T is a complex bi-fundamental gauge field so that
1521: $D_{\mu}T=\partial_{\mu}T-iTA_{\mu}+i{\overline{A}}_{\mu}T$.
1522: Dimensional reduction of this action should give the correct action
1523: for the dynamics of lower dimensional D$p$-${\overline{{\rm D}p}}$
1524: systems including our $p=0$ case.  As usual, we substitute
1525: ${\overline{A}}_{i}=(2\pi\alpha^{\prime})^{-1}{\overline{X}}_{i}\,,\,
1526: A_{i}=(2\pi\alpha^{\prime})^{-1}X_{i}$.  Performing field
1527: redefinitions, a nonabelian generalization of this toy approximate
1528: action is
1529: %
1530: \begin{eqnarray}\label{nonabeliantoy}
1531: S&=&-\int dx^0 {\rm{Tr}}\
1532: \exp\left(-\frac{\pi^2\alpha'}{4\ln2}T^{\dagger}T\right)
1533: \biggl[
1534: \frac{1}{2g_{\rm YM}^2}(\partial_{0}T^{\dagger}\partial^{0}T
1535: +i\partial_{0}T^{\dagger}({\overline{A}}^{0}T-TA^{0})  \nonumber\\
1536: %
1537: && +i(A^{0}T^{\dagger}-T^{\dagger}{\overline{A}}^{0})\partial_{0}T
1538: -(A^{0}T^{\dagger}-T^{\dagger}{\overline{A}}^{0})
1539: ({\overline{A}}_{0}T-TA_{0}) \nonumber\\
1540: %
1541: &&
1542: -(X^{i}T^{\dagger}-T^{\dagger}{\overline{X}}^{i})
1543: ({\overline{X}}_{i}T-TX_{i}))
1544: +2\tau_{0}\mathbb{I}_{N\times N}\biggr] \nonumber\\
1545: %
1546: && + S'_{\rm D0}[X^i,A^0]+S'_{\overline{\rm D0}}
1547: [{\overline{X}}^i,{\overline{A}}^0] \,,
1548: \end{eqnarray}
1549: %
1550: where we use the usual definition of the 't Hooft coupling, $1/g_{\rm
1551: YM}^2=4\pi^2 g_{s}^{-1}\ell_s^{3}$, and where $S'_{\rm D0}$ and
1552: $S'_{\overline{\rm D0}}$ are the corresponding low-energy actions for
1553: D0 and $\overline{\rm D0}$ respectively.  Obviously, $T$ is in the
1554: $N\times {\overline{N}}$ bifundamental representation.
1555: 
1556: Now, in order to be able to compute the thermal partition
1557: function, we need to Euclideanize the Minkowskian action. First,
1558: we analytically continue the timelike component of both $U(N)$
1559: gauge fields as well as Minkowskian time direction $x^0$ while
1560: leaving the other fields untouched,
1561: %
1562: \begin{equation}
1563: x^0=-i\tau \,,\quad
1564: %
1565: iS_E=S_M \,,\quad iA^{0}_E=A^{0}_M \,,\quad
1566: %
1567: i{\overline{A}}^{0}_E={\overline{A}}^{0}_M \,.
1568: \end{equation}
1569: %
1570: Next, we Fourier expand $T$, $A^{0}$, ${\overline{A}}^{0}$, $X^{i}$
1571: and ${\overline{X}}^{i}$ in $\tau$ with a periodicity $\beta$ given by
1572: the the inverse temperature.  Since all these variables are bosonic,
1573: they have periodic boundary conditions,
1574: %
1575: \begin{equation}
1576: \left\{ A_{0}(\tau) , {\overline{A}}_{0}(\tau), X^i(\tau) ,
1577: {\overline{X}}^i(\tau) , T(\tau) \right\}
1578: =\frac{1}{\sqrt{\beta}}\sum_{\ell\in\mathbb{Z}} \left\{ A_{0\ell} ,
1579: {\overline{A}}_{0\ell} , X^i_{\ell} , {\overline{X}}^i_{\ell} , T_\ell
1580: \right\} e^{\frac{-2\pi i \ell}{\beta}\tau} \,.
1581: \end{equation}
1582: %
1583: Writing the action in terms of Fourier modes is straightforward.  We
1584: take the above expansions and perform the Euclidean time integral,
1585: using the following {\em variational} correlators in Wick contractions
1586: of Feynman diagrams contributing to the free energy (in our case, the
1587: amplitude $\langle S_{II E}-S_{0 E}\rangle_{0}$,)
1588: %
1589: \begin{eqnarray}
1590: \langle X^{i}_{\ell AB} X^{j}_{m CD}\rangle_{0} &=&\sigma_{\alpha
1591: \ell}^2 \delta^{i j}\delta_{\ell+m}\delta_{AD}\delta_{BC}\quad\quad
1592: i,j=1,2 \,,\nonumber\\ 
1593: %
1594: \langle X^{a}_{\ell AB} X^{b}_{m CD}\rangle_{0} &=&\Delta_{\alpha l}^2
1595: \delta^{ab}\delta_{\ell+m}\delta_{AD}\delta_{BC}\quad\quad a,b=3..9
1596: \,,\nonumber\\
1597: %
1598: \langle A_{00AB} A_{00CD}\rangle_{0} &=&
1599: \rho_{0}^{2} \delta_{AD}\delta_{BC} \,,\nonumber\\
1600: %
1601: \langle T^{\dagger}_{\ell AB} T_{m
1602: CD}\rangle_{0}&=&\xi_{\ell}^2\delta_{\ell m}\delta_{AD}\delta_{BC}
1603: \,.
1604: \end{eqnarray}
1605: %
1606: and similarly for the barred variables.  The need for separation of
1607: transverse scalar directions comes from the limitations of the
1608: original superspace formulation \cite{kabat2} of the D0-brane
1609: supersymmetric matrix quantum mechanics. Namely, for each copy of the
1610: QM, there are 2 scalars $X^{i}$ in the gauge multiplet and $7$ other
1611: $\phi^{a}$ scalars coming from scalar multiplet.  In other words, the
1612: original $SO(9)$ $R$-symmetry is broken to $SO(2) \times\ G_2$
1613: ($\phi^{a}$ are a $\textbf{7}$ of $G_{2}$).
1614: 
1615: The numerical simulation of this action, which is an approximation
1616: to the system of D0-branes, anti-D0-branes and the tachyon sector
1617: where both copies of the D0-brane theory are supersymmetric is
1618: computationally very expensive, partly because of the need to find
1619: the solution to hundreds of nonlinear equations. In addition, we
1620: have to cover mapping out of the phase portrait as a function of
1621: $\beta$ and $g_sN$.  We therefore report on some initial results
1622: in a more stripped-down version of our model.
1623: 
1624: %====================================================================+
1625: \subsection{The toy model}
1626: %
1627: In this section, we look at a ``toy'' model which consists of
1628: large-$N$ bosonic matrix quantum mechanics (representing the D0-brane
1629: and anti-D0-brane theories) plus our bi-fundamental charged tachyon.
1630: 
1631: This toy model is obviously a good approximation to the original
1632: problem in the high-temperature limit where the Euclidean time circle
1633: is so small that nonzero thermal KK modes are heavy. In this limit,
1634: there is no antiperiodic fermion left after reduction to $0+0$
1635: dimensions. Of course, we are interested in low temperatures where the
1636: supergravity description is valid and quantum mechanics is strongly
1637: coupled.
1638: 
1639: There are reasons to believe why this toy model could demonstrate some
1640: generic behaviors of interest to us, which will continue to be true
1641: even in the case where we have two copies of {\em supersymmetric}
1642: matrix quantum mechanics. It is well known \cite{Polchinski} that the
1643: extent of the ground state of a system of D0-branes, defined in terms
1644: of fluctuations of transverse scalars, is controlled by the 't Hooft
1645: coupling. It is intuitively plausible to expect that, at strong
1646: coupling, the tachyon could acquire a large induced mass through its
1647: coupling to these large transverse scalar fluctuations.  As has been
1648: observed in \cite{kabat1}, both bosonic matrix quantum mechanics and
1649: its supersymmetric version have large transverse fluctuations.  So if
1650: one takes this intuition seriously, as far as the dynamics of the
1651: tachyon is concerned in the low energy approximation, even the bosonic
1652: version could be a good approximation to the right
1653: physics\footnote{Although irrelevant to the above discussion, the
1654: difference is in the actual value of the ground state energy, which is
1655: -- for each copy -- of course zero in the supersymmetric case but
1656: large and positive in the bosonic case.}
1657: 
1658: In the model of Danielsson et al \cite{Danielsson} which motivated our
1659: work here, the dynamical assumption is that the tachyon field is
1660: stabilized by finite-temperature strong-coupling effects about $T=0$.
1661: We therefore expand $T= 0 + t$, where $t$ is the fluctuation part, and
1662: plan to show the consistency of this assumption {\em a posteriori} by
1663: showing that the mass is indeed positive -- and large\footnote{We have
1664: ignored the higher powers in the expansion of the exponential
1665: prefactor in (\ref{nonabeliantoy}) sitting in front of the Lagrangian;
1666: this will be justified {\rm a posteriori} by the large-mass finding.}
1667: -- at strong coupling.  Therefore, the terms relevant to our
1668: discussion which are quadratic in the tachyon fluctuations $t$ can be
1669: written
1670: %
1671: \begin{eqnarray}
1672: S_{E}&=&\frac{1}{g^2_{\rm YM}}\int_{0}^{\beta}
1673: d\tau(\frac{1}{2}{\rm{Tr}} D_{\tau}X^{i}D_{\tau}X^{i} -\frac{1}{4}
1674: {\rm{Tr}}[X^{i},X^{j}][X^{i},X^{j}])\nonumber \\
1675: %
1676: && +\frac{1}{g^2_{\rm YM}}\int_{0}^{\beta} d\tau(\frac{1}{2}{\rm{Tr}}
1677: D_{\tau}{\overline{X}}^{i}D_{\tau}{\overline{X}}^{i} -
1678: \frac{1}{4}{\rm{Tr}}[{\overline{X}}^{i},{\overline{X}}^{j}]
1679: [{\overline{X}}^{i},{\overline{X}}^{j}]) \nonumber\\
1680: %
1681: && +\int_{0}^{\beta} d\tau {\rm{Tr}}[\frac{1}{2g^2_{\rm
1682: YM}}
1683: \biggl(\partial_{\tau}t^{\dagger}\partial_{\tau}t
1684: +A^{0}t^{\dagger}tA^{0}+t^{\dagger}{\overline{A}}^{0}
1685: {\overline{A}}^{0}t +X^{i}t^{\dagger}tX^{i} +
1686: t^{\dagger}{\overline{X}}^{i}{\overline{X}}^{i}t
1687: \nonumber\\ 
1688: &&
1689: %
1690: - \partial_0 t^\dagger ({\overline{A}}^0 t - t A^0) + (A^0
1691: t^\dagger-t^\dagger {\overline{A}}^0) \partial_0 t \biggr)
1692: +2\tau_{0}\mathbb{I}_{N\times
1693: N}-\frac{2\tau_0\pi^2\alpha'}{4\ln2}t^{\dagger}t\biggr] \,.
1694: \end{eqnarray}
1695: %
1696: Morally, we are going to treat the tachyon as background.  By this we
1697: simply mean that we are interested in finding the free energy as a
1698: function[al] of the tachyon, after integrating out the other fields in
1699: the problem -- nonperturbatively using VPT.  For this reason, no trial
1700: action for the tachyon is introduced; the VPT techniques are used for
1701: the fields other than the tachyon.
1702: 
1703: Using VPT, we have
1704: %
1705: \begin{equation}
1706: \beta F=\beta F_{0} + \beta {\overline{F}}_{0}+\langle S_{\rm
1707: D0}+S_{\overline{\rm D0}} +
1708: S_{T}-S_{0}\rangle_{0}-\frac{1}{2}\langle(S_{III})^2\rangle_{0,C}
1709: \quad (+\ldots) \,,
1710: \end{equation}
1711: %
1712: where $S_T$ is the tachyon contribution to the action, and the
1713: subscript zero refers to the contribution coming from the trial free
1714: action for the other fields.
1715: 
1716: As a first step, we can neglect the terms in the action that are cubic
1717: in the tachyon, as far as the VPT expansion is concerned.  There is a
1718: good reason why this is justified.  Namely, because the cubic
1719: contribution to the free energy enters at the order of
1720: $\mathcal{O}(t^4)$, which is small at small-$t$. 
1721: 
1722: Following the story for a single clump of D0-branes \cite{kabat1}, we
1723: put in the ghost field from gauge-fixing, whose Fourier modes are
1724: denoted $s_{\ell\not=0}$ (obviously, for physical reasons there is no
1725: zero mode).  We also use equations from pages 25-27 of that work; to
1726: save space here we only write here what changes when we introduce our
1727: tachyon field.  
1728: 
1729: We also introduce convenient dimensionless units. These are defined
1730: such that in term of them one gets an overall factor of $N^2$ in the
1731: free energy, and the rest is just a function of dimensionless
1732: quantities and $g_{s}N$.  We define newly dimensionless variables 
1733: with tildes; we have
1734: %
1735: \begin{equation}
1736: \tilde{\beta}=\beta {f^{1/3}} \,,\quad
1737: \tilde{\rho}_{0}^2=\frac{N\rho_{0}^2}{f^{1/3}} \,,\quad
1738: \tilde{\sigma}_{l}^2=\frac{N\sigma_{l}^2}{f^{1/3}} \,,\quad
1739: {\tilde{t}}_\ell^\dagger {\tilde{t}}_\ell=\frac{N t^\dagger_\ell
1740: t_\ell}{f^{1/3}} \,,
1741: \end{equation}
1742: %
1743: and similarly for the barred variables.  Here $f=(g_{\rm YM}^2
1744: N)^{1/3}$.  Note that in this bosonic model we do not have to treat
1745: the $X^i,{\overline{X}}^i$ fields for $i=1,2$ differently than for
1746: $i=3\ldots 9$ because there is no need to worry about maintaining
1747: explicit supersymmetry; therefore we do not have $\Delta_\ell$ and we
1748: just use $\sigma_\ell$ to represent correlators of all nine position
1749: fields.  Note that we are measuring the tachyon expectation value in
1750: string units.  
1751: 
1752: Henceforth, we drop the tildes for notational simplicity.  Then we
1753: obtain the following expression for the free energy as a function of
1754: the tachyon
1755: %
1756: \begin{eqnarray}
1757: \beta F&=&\beta F(\lambda)_{\square}+\beta
1758: F({\overline{\lambda}})_{\square} +
1759: \frac{N}{\lambda}\langle
1760: {\rm{Tr}}(U+U^{\dagger})\rangle_{\square}+
1761: \frac{N}{{\overline{\lambda}}}\langle
1762: {\rm{Tr}}({\overline{U}}+{\overline{U}}^{\dagger})\rangle_{\square}
1763: \nonumber \\
1764: %
1765: &&
1766: -\frac{9N^2}{2}\sum_{\ell}\log\sigma_\ell^2
1767: -\frac{9N^2}{2}\sum_{\ell}\log{\overline{\sigma}}_\ell^{2}
1768: +N^2\sum_{\ell\neq 0}\log s_{\ell}
1769: +N^2\sum_{\ell\neq 0}\log {\overline{s}}_{\ell}\nonumber\\  &&+
1770: %
1771: \frac{9N^2}{2}\sum_{\ell}((\frac{2\pi l}{\beta})^2
1772: \sigma_{\ell}^{2}-1) +\frac{9N^2}{2}\sum_{\ell}((\frac{2\pi
1773: l}{\beta})^2{\overline{\sigma}}_{\ell}^{2}-1) \nonumber\\
1774: %
1775: && +N^2\sum_{\ell\neq 0}((\frac{2\pi
1776: l}{\beta})^2s_{\ell}+1)+N^2\sum_{\ell\neq 0}((\frac{2\pi
1777: l}{\beta})^2{\overline{s}}_{\ell}+1) +
1778: \frac{9N^2}{\beta}\rho_{0}^{2}\sum_{\ell}\sigma_{\ell}^{2} \nonumber\\
1779: &&
1780: %
1781: + \frac{9N^2}{\beta}{\overline{\rho}}_{0}^{2}
1782: \sum_{\ell}{\overline{\sigma}}_{\ell}^{2}
1783: +\frac{36N^2}{\beta}(\sum_{\ell}\sigma_{\ell}^{2})^2 +
1784: \frac{36N^2}{\beta}(\sum_{\ell}{\overline{\sigma}}_{\ell}^{2})^2
1785: \nonumber\\
1786: %
1787: && +N^2\sum_{\ell}\frac{1}{2}(\frac{2\pi l}{\beta})^2
1788: {\rm{Tr}}\frac{t^{\dagger}_{\ell}t_{\ell}}{N^2}+
1789: \frac{N^2}{2\beta}(\rho^{2}_{0} + {\overline{\rho}}^{2}_{0})
1790: \sum_{\ell}{\rm{Tr}}\frac{t^{\dagger}_{\ell}t_{\ell}}{N^2} \nonumber\\
1791: %
1792: &&+ \frac{9N^2}{2\beta}(\sum_{\ell}\sigma^{2}_{\ell} +
1793: \sum_{\ell}{\overline{\sigma}}^{2}_{\ell})
1794: \sum_{\ell}{\rm{Tr}}\frac{t^{\dagger}_{\ell}t_{\ell}}{N^2}
1795: \nonumber\\ &&
1796: %
1797: -\frac{(2\pi)^{4/3}N^2}{8\ln2(g_{s}N)^{2/3}}\sum_{\ell}
1798: {\rm{Tr}}\frac{t^{\dagger}_{\ell}t_{\ell}}{N^2} +
1799: \frac{2(2\pi)^{2/3}N^2\beta}{(g_{s}N)^{4/3}} \,.
1800: \end{eqnarray}
1801: %
1802: 
1803: In what follows, we use a natural ansatz for the propagators
1804: %
1805: \begin{equation}
1806: \sigma^{2}_{\ell}= \left( {(\frac{2\pi \ell}{\beta})^2+m^{2}}
1807: \right)^{-1} \,,
1808: \end{equation}
1809: %
1810: and similarly for the barred variables.
1811: 
1812: The gap equations are easily obtained. We have (and similarly for
1813: barred variables)
1814: %
1815: \begin{equation}
1816: m^{2}=\frac{2\rho^{2}_{0}}{\beta} +
1817: \frac{16}{\beta}\sum_{\ell}\sigma^{2}_{\ell} +
1818: \frac{1}{\beta}\sum_{\ell}\frac{t^{\dagger}_{\ell}t_{\ell}}{N^2} \,,
1819: \end{equation}
1820: %
1821: where $\rho_0$ can be explicitly obtained in terms of $\lambda$, the
1822: coupling in the one-plaquette action, because the one-plaquette theory
1823: is soluble in the large-$N$ limit.
1824: 
1825: For $\lambda\leq 2$,
1826: %
1827: \begin{equation}
1828: -1+\frac{2}{\beta}\left(\frac{18}{\beta}\sum_{\ell}\sigma^{2}_\ell +
1829: \frac{1}{\beta}\sum_{\ell}{\rm{Tr}}
1830: \frac{t^{\dagger}_{\ell}t_{\ell}}{N^2}\right)
1831: \left[1-(1-\frac{2}{\lambda})\log(1-\frac{\lambda}{2})\right]=0 \,,
1832: \end{equation}
1833: %
1834: while for $\lambda\geq 2$, 
1835: %
1836: \begin{equation}
1837: \lambda =\beta/\left(\frac{18}{\beta}\sum_{\ell}\sigma^{2}_\ell +
1838: \frac{1}{\beta}
1839: \sum_{\ell}{\rm{Tr}}\frac{t^{\dagger}_{\ell}t_{\ell}}{N^2}\right) \,.
1840: \end{equation}
1841: %
1842: We will solve this system of highly nonlinear, coupled equations.  The
1843: strategy will be to treat the tachyon as a background.  Solving the
1844: gap equations, and substituting back the solution as a function of the
1845: tachyon, then gives the free energy as a function of the (background)
1846: tachyon.  Given this effective action for the tachyon, it is then
1847: straightforward to read off the tachyon effective mass. The plan will
1848: be to work out the two-dimensional ``phase portrait'' of the system
1849: (with coordinates $g_{s}N $ and $\beta$), and find out whether any
1850: tachyon stabilization is possible. By this we simply mean that there
1851: are points in this two-dimensional portrait where the effective
1852: tachyon mass is positive.
1853: 
1854: Before we move to displaying our numerical results, a number of
1855: remarks are in order.
1856: 
1857: \begin{itemize}
1858: 
1859: \item As we saw earlier, in the supergravity limit in which we are
1860: interested, $g_{s}N \gg 1$.  By inspection of the equations it might
1861: then seem that, in the units we use, the wrong-sign mass term for the
1862: tachyon is suppressed and we are done!  This quick-and-dirty reasoning
1863: is, however, too na\"ive. In fact, in order not to excite any massive
1864: open string state, the temperature must be small in string units.  In
1865: turn, this implies that
1866: %
1867: \begin{equation}
1868: \beta \gg (g_{s}N)^{1/3} \gg 1 \,,
1869: \end{equation}
1870: %
1871: where $\beta$ is the dimensionless inverse temperature. With such low
1872: dimensionless temperatures, it becomes more and more difficult to
1873: overcome the wrong-sign mass term through the tachyon couplings to the
1874: transverse coordinates $X^{i},{\overline{X}}^i$ and the world-volume
1875: gauge fields $A_{0},{\overline{A}}_0$. 
1876: 
1877: \item It is crucial to remember that all variational parameters are
1878: {\em implicit} functions of the tachyon $t$.  So it is certainly not
1879: true that we can simply read off the effective tachyon mass from the
1880: the coefficients of the quadratic tachyon fluctuation in the
1881: expression for the free energy.  For example, terms like
1882: $(N^2/\beta)\rho^{2}_{0}\sum_{\ell}\sigma^{2}_{\ell}$, which do not
1883: have any {\em explicit} background tachyon dependence, are becoming
1884: tachyon-dependent through the gap equations.  All such terms make
1885: contributions to the effective tachyon mass, as one expands the
1886: variational parameters around $\sum_{\ell}{\rm{Tr}}
1887: t^{\dagger}_{\ell}t_{\ell}=0$. (Note that these expansions exist,
1888: since the solution to the gap equations is differentiable with respect
1889: to $t$ in the vicinity of $\sum_{\ell}{\rm{Tr}}
1890: t^{\dagger}_{\ell}t_{\ell}=0$.)
1891: 
1892: \item Obviously, it is always possible to stabilize the tachyon at
1893: weak couplings, but the temperatures needed for this are above the
1894: Hagedorn temperature. It is easy to see this in perturbation theory,
1895: partly because at such high temperatures the D0 QM is perturbative.
1896: Perturbatively, to stabilize the tachyon, it is sufficient to have
1897: %
1898: \begin{equation}
1899: T>\frac{1}{(g_{s}N)^{4/3}\ell_s} \gg \frac{1}{\ell_s} \,,
1900: \end{equation}
1901: %
1902: for small $g_{s}N$.
1903: 
1904: \end{itemize}
1905: 
1906: In our investigations, we use the Metropolis algorithm.  This is a
1907: fast method generally; it really comes into its own for numerical
1908: investigations of the supersymmetric model, the numerical results from
1909: which we hope to report on in the near future.  We summarize some
1910: important features of the Metropolis algorithm in the Appendix.
1911: 
1912: %====================================================================+
1913: \subsection{Numerical results}
1914: %
1915: We are interested in the sign of the mass of the tachyon zero mode,
1916: for each point in the two-dimensional parameter space, coordinatized
1917: by $g_sN$ and $\beta$ which is the dimensionless inverse temperature
1918: in 't Hooft units.
1919: 
1920: We now plot our results for the phase portrait of the toy model
1921: theory.  Inverse temperature, in dimensionless 't Hooft units, is on
1922: the $y$-axis, while the open-string coupling $g_s N$ on the $x$-axis.
1923: (As a reminder, the parameter $g_s N$ tells us the bare negative
1924: mass-squared of the tachyon.)  Points at which the tachyon mass
1925: squared is positive are indicated by a cross, and those at which it is
1926: negative are indicated by a circle.  The green curve delineates where
1927: the Hagedorn phenomenon is important; points above the curve are below
1928: the Hagedorn temperature.  The red straight line delineates the region
1929: where the gauged QM theory has a strong effective coupling (set by the
1930: inverse temperature and 't Hooft coupling); the temperatures where the
1931: quantum mechanical theory is strongly coupled lie above the red line.
1932: 
1933: For clarity, we separate the regimes of smaller $g_s N $ from the
1934: regime of larger $g_s N $.  
1935: 
1936: \begin{figure}
1937: \begin{center}
1938: \epsfysize=3truein
1939: \epsffile{weakresults.eps}
1940: \end{center}
1941: \caption{Plot of results for the $g_sN\leq 2$ section of the phase
1942: portrait.  }
1943: \label{fig:fig1}
1944: \end{figure}
1945: 
1946: \begin{figure}
1947: \begin{center}
1948: \epsfysize=3truein \epsffile{strongresults.eps}
1949: \end{center}
1950: \caption{Plot of results for the $g_sN \geq 2$ section of the phase
1951: portrait.  }
1952: \label{fig:fig2}
1953: \end{figure}
1954: 
1955: As we expected for tachyon stabilization, at small $g_sN$, one has to
1956: go to very high temperatures; this is clear from Figure
1957: \ref{fig:fig1}.  As we see for large $g_sN$, on the other hand, even
1958: for low temperatures compared to the Hagedorn temperature, there is a
1959: possibility of tachyon stabilization.
1960: 
1961: %====================================================================+
1962: \section{Discussion}
1963: 
1964: In this paper, following the lines of \cite{Danielsson}, we have
1965: examined the validity of the brane-antibrane model for neutral black
1966: branes of various dimensionality.  What we have found is that this
1967: D$p$-${\overline{{\rm D}p}}$ picture works for arbitrary
1968: $p$. Interestingly, this is true even for the ``dilatonic'' branes,
1969: where the supergravity entropy cannot be written in terms of the
1970: entropy of a free world-volume gauge theory.  In the non-conformal
1971: cases, using IMSY duality proves to be working, relating strongly
1972: coupled gauge theories to the corresponding near-extremal backgrounds.
1973: We were able to show that, even after adding general angular momenta
1974: quantum numbers, the brane-antibrane model is able to reproduce the
1975: supergravity entropy.  The exact agreement in the entropy between the
1976: microscopic brane-antibrane side and the supergravity side comes at
1977: the expense of a renormalization of the mass by a $p$-dependent power
1978: of two, which we interpret as a binding energy, and a renormalization
1979: of the angular momenta by a simple factor of two.  We also find that
1980: the extent of the wavefunction of the $Dp-\overline{Dp}$ system
1981: exhibits the same scaling with the mass as does the radius of the
1982: horizon of the corresponding black brane.
1983: 
1984: In the original proposal of \cite{Danielsson}, the authors argued that
1985: the tachyon stabilization at finite temperature and strong coupling
1986: could lead to a reappearance of open string degrees of freedom.  Even
1987: at the field theory level, questions regarding strong coupling
1988: dynamics are often hard to answer, but in the special case of
1989: D0-${\overline{{\rm D}0}}$ there is a possibility to address some of
1990: the aspects in the context of a simple toy model following the lines
1991: of \cite{kabat1,kabat2}.  Using techniques developed there, we found
1992: the effective action for the tachyon numerically, in the context of
1993: the toy model. We have seen signals of tachyon stabilization in this
1994: toy model, at strong open string couplings and low temperatures
1995: compared to the string scale.  This might open a window to see clearly
1996: why it is possible to have a long-lived state of D0-branes and
1997: $\overline{{\rm D}0}$-branes, without decaying into closed strings at
1998: couplings of order one and low temperatures.
1999: 
2000: Our investigation here was for the case of neutral black branes, where
2001: the numbers of microscopic constituent branes $N$ and antibranes
2002: ${\overline{N}}$ are equal.  A natural extension of our work here will
2003: be to consider the case where the $D=10$ black branes carry $p$-brane
2004: charge $Q_p=N-{\overline{N}}$.  We hope to report on this story in the
2005: future\footnote{The $p=3$ case with charge was of course previously
2006: considered in \cite{Danielsson,Guijosa}.  It will be interesting to
2007: check whether this agreement holds up also in the {\em non}-conformal
2008: cases.  We thank Alberto Guijosa for a discussion in this regard.}.
2009: 
2010: Inspired by numerical investigations, we might speculate about the
2011: possible shape of the potential\footnote{at least for $|T|$ in the
2012: vicinity on $T=0$} in this model, say for the simplest case of a
2013: diagonal tachyon condensate at finite temperatures
2014: %
2015: \begin{equation}
2016: \beta F \propto e^{-\gamma |T|^2}(1+\lambda(N,g_{YM},\beta)
2017: |T|^2+\ldots)
2018: \end{equation}
2019: %
2020: where $\lambda > \gamma$. This potential has one minimum and two
2021: symmetrically located maxima. The minimum would correspond to the
2022: tachyon stabilization, i.e. the tachyon particle is classically
2023: stable\footnote{The kinetic term probably is not canonically
2024: normalized in the original field.}, even though this vacuum would be
2025: unstable due to quantum and thermal fluctuations and the finiteness of
2026: the height of the barrier. The tachyon can be expected to decay
2027: eventually into closed string radiation. In other words, this
2028: temporary and approximate decoupling between the open string modes and
2029: the closed strings degrees of freedom -- which has apparently
2030: manifested itself in our ability to compute the entropy of the
2031: closed-string background out of Yang-Mills degrees of freedom in the
2032: context of brane-antibrane model -- would not last forever, and the
2033: tachyon particle ground state would have a finite lifetime. If this
2034: picture corresponds well to the true stringy physics, then it would be
2035: plausible to think of the unstable maxima discussed above as being the
2036: unstable state of a D0-$\overline{{\rm D}0}$ pair created from the
2037: energy in the gas living on the worldvolume of the clump of D0s and
2038: $\overline{{\rm D}0}$s, due to the quantum and thermal fluctuations.
2039: This interpretation also leads to an estimate for the
2040: lifetime. Climbing up the hill to create such a pair requires energies
2041: at least of the order of $1/g_s$, so the whole tunnelling process is
2042: suppressed by $e^{-1/g_{s}}$.
2043: 
2044: %====================================================================+
2045: %\newpage
2046: \section*{Acknowledgements}
2047: 
2048: We would like to thank Joel Giedt, Martin Kruczenski, Erich Poppitz,
2049: and Lenny Susskind for helpful discussions.  We would especially like
2050: to thank Dan Kabat for substantive helpful conversations, particularly
2051: about the numerical collaborations \cite{kabat1,kabat2}.  We also
2052: thank Kaveh Khodjasteh and Johannes Martin for assistance with
2053: computer resources.
2054:  
2055: %====================================================================+
2056: \section*{Appendix}
2057: 
2058: The problem of finding the common solution to a set of nonlinear
2059: coupled algebraic equation can be though of a minimization process.
2060: To see this, consider the collection of the equations to be solved,
2061: $\mathcal{G}=\{f_{1}=0,...,f_{n}=0 \}$.  Clearly, a quadratic form
2062: $\mathcal{H}$ can be made out of this, viz.  $ \mathcal{H}=\sum_{i}
2063: f^2_{i} $.  If there is a solution to $\mathcal{G}$, this solution
2064: would be the global minimum of $\mathcal{H}$ i.e.,$
2065: f_{1}=0,...,f_{n}=0 $.
2066: 
2067: To solve this nonlinear system of equations, we use the Monte Carlo
2068: local update method (``Metropolis'' algorithm). This method is one of
2069: the stochastic search methods widely used for minimization of
2070: multivariable functions.
2071: 
2072: This method does not get trapped in local minima, because of a smart
2073: choice of control variables. The way this method accomplishes the goal
2074: is that, by introducing an unphysical temperature, thermal
2075: fluctuations of the configuration vector are allowed to happen in all
2076: possible directions in the configuration space - even in the ``wrong
2077: directions'' (which increase the Hamiltonian).  By decreasing the
2078: temperature very slowly, one lets the vector find the global minimum
2079: (or minima). For a vector trapped in a local minimum, there would be a
2080: certain probability to climb up the barrier because of the thermal
2081: fluctuations, and so the vector could find its way all the way down to
2082: the global minimum subject to a large number of iterations.  This
2083: story is actually very similar to the physics of the annealing
2084: process, and it is sometimes called the ``stimulated annealing''
2085: method.  With this method, a vector stores the starting point
2086: configuration. This configuration is changed randomly to generate new
2087: child configurations. A new configuration gets accepted with the
2088: probability $1$ if it has a smaller Hamiltonian than its parent, but
2089: if it has not then it would be accepted with the following
2090: Boltzmann-like probability $ P_{\rm acceptance}\sim e^{-\beta_{\rm
2091: fake}\Delta \mathcal{H}}$, where $\beta_{\rm fake}$ is a unphysical
2092: parameter acting like an inverse temperature, and $\Delta \mathcal{H}$
2093: is the energy difference between the child and parent
2094: configurations. After many iterations, and at the same time lowering
2095: the temperature slowly, the configuration vector stabilizes on the
2096: one(s) which minimize(s) the $\mathcal{H}$.
2097: 
2098: The Metropolis algorithm has strong advantages over the Newton-Raphson
2099: (NR) method, especially when the dimensionality of the configuration
2100: space is big.  Under such circumstances, the NR method becomes very
2101: inefficient, and its starting point dependence can be prohibitively
2102: large.  Practically, this makes it very hard to reach the solution in
2103: reasonable computer time.  The great advantage of the Metropolis
2104: method is the fact that it is almost independent of starting
2105: point.\footnote{Of course, any configuration space could be composed
2106: of disconnected pieces, such that one can not get to any arbitrary
2107: point by starting from another arbitrary point and moving through the
2108: potential.}  Another advantage of Metropolis is that it naturally
2109: avoids getting stuck in local minima, a problem to which NR often
2110: succumbs.
2111: 
2112: %====================================================================+
2113: %\newpage
2114: 
2115: \addcontentsline{toc}{section}{References}
2116: 
2117: \begin{thebibliography}{99}
2118: 
2119: \bibitem{Danielsson} Ulf H. Danielsson, Alberto Guijosa, Martin
2120: Kruczenski, ``Brane Anti-Brane Systems at Finite Temperature and the
2121: Entropy of Black Branes, JHEP 0109:011,2001 [arXiv:hep-th/0106201];
2122: U. H. Danielsson, A. Guijosa and M.~Kruczenski, Black brane entropy
2123: from brane-antibrane systems, arXiv:gr-qc/0204010.
2124: 
2125: \bibitem{kabat1}D. Kabat and G. Lifschytz,
2126: Approximations for strongly-coupled supersymmetric quantum mechanics,
2127: Nucl. Phys. B {\bf 571}, 419 (2000)
2128: [arXiv:hep-th/9910001].
2129: 
2130: \bibitem{kabat2}
2131: D. Kabat, G. Lifschytz and D. A. Lowe,
2132: Black hole thermodynamics from calculations in strongly coupled
2133: gauge  theory,
2134: Phys. Rev. Lett.  {\bf 86}, 1426 (2001)
2135: [Int. J. Mod. Phys. A {\bf 16}, 856 (2001)
2136: [arXiv:hep-th/0007051].
2137: 
2138: \bibitem{BFSS} T.~Banks, W.~Fischler, S.~H.~Shenker and L.~Susskind,
2139: ``M theory as a matrix model: A conjecture,'' Phys.\ Rev.\ D {\bf 55},
2140: 5112 (1997) [arXiv:hep-th/9610043].
2141: 
2142: \bibitem{KlebanovSusskind} I.~R.~Klebanov and L.~Susskind,
2143: ``Schwarzschild black holes in various dimensions from matrix
2144: theory,'' Phys.\ Lett.\ B {\bf 416}, 62 (1998) [arXiv:hep-th/9709108].
2145: 
2146: \bibitem{peetCQGreview} A.~W.~Peet, ``The Bekenstein formula and
2147: string theory (N-brane theory),'' Class.\ Quant.\ Grav.\ {\bf 15},
2148: 3291 (1998) [arXiv:hep-th/9712253].
2149: 
2150: \bibitem{Englert} F.~Englert and E.~Rabinovici, ``Statistical entropy
2151: of Schwarzschild black holes,'' Phys.\ Lett.\ B {\bf 426}, 269 (1998)
2152: [arXiv:hep-th/9801048].
2153: 
2154: \bibitem{dasetal} S.~Das, A.~Ghosh and P.~Mitra, ``Statistical entropy
2155: of Schwarzschild black strings and black holes,'' Phys.\ Rev.\ D {\bf
2156: 63}, 024023 (2001) [arXiv:hep-th/0005108].
2157: 
2158: \bibitem{itzhaki}
2159: N. Itzhaki, J. M. Maldacena, J. Sonnenschein and S. Yankielowicz,
2160: Supergravity and the large N limit of theories with sixteen
2161: supercharges,
2162: Phys.\ Rev.\ D {\bf 58}, 046004 (1998)
2163: [arXiv:hep-th/9802042].
2164: 
2165: \bibitem{KlebanovTseytlin}
2166: I.~R.~Klebanov and A.~A.~Tseytlin, ``Entropy of Near-Extremal
2167: Black p-branes,'' Nucl.\ Phys.\ B {\bf 475}, 164 (1996)
2168: [arXiv:hep-th/9604089].
2169: 
2170: \bibitem{Maldacena1}G. T. Horowitz, J. M. Maldacena and A. Strominger,
2171: Nonextremal Black Hole Microstates and U-duality, Phys. Lett. B {\bf
2172: 383}, 151 (1996) [arXiv:hep-th/9603109].
2173: 
2174: \bibitem{peet1} S. S. Gubser, I. R. Klebanov and A. W. Peet, Entropy
2175: and Temperature of Black 3-Branes, Phys.\ Rev.\ D {\bf 54}, 3915
2176: (1996) [arXiv:hep-th/9602135].
2177: 
2178: \bibitem{HarmarkObers}
2179: T.~Harmark and N.~A.~Obers,
2180: ``Thermodynamics of spinning branes and their dual field theories,''
2181: JHEP {\bf 0001}, 008 (2000)
2182: [arXiv:hep-th/9910036].
2183: 
2184: \bibitem{peet}
2185: A. W. Peet and J. Polchinski
2186: UV/IR relations in AdS dynamics,
2187: Phys.\ Rev.\ D {\bf 59}, 065011 (1999)
2188: [arXiv:hep-th/9809022].
2189: 
2190: \bibitem{KabatLowe}
2191: N.~Iizuka, D.~Kabat, G.~Lifschytz and D.~A.~Lowe, ``Probing black
2192: holes in non-perturbative gauge theory,'' Phys.\ Rev.\ D {\bf 65},
2193: 024012 (2002) [arXiv:hep-th/0108006].
2194: 
2195: \bibitem{kabat3}
2196: D. Kabat and G. Lifschytz,
2197: Tachyons and black hole horizons in gauge theory,
2198: JHEP {\bf 9812}, 002 (1998)
2199: [arXiv:hep-th/9806214].
2200: 
2201: \bibitem{Gross}
2202: D. J. Gross and E. Witten,
2203: Possible Third Order Phase Transition In The Large N Lattice
2204: Gauge Theory, Phys. Rev. D {\bf 21}, 446 (1980).
2205: 
2206: \bibitem{Kraus} P. Kraus and F. Larsen, Boundary string field theory
2207: of the DD-bar system, Phys. Rev. D {\bf 63}, 106004 (2001)
2208: [arXiv:hep-th/0012198].
2209: 
2210: \bibitem{Tadashi} T. Takayanagi, S. Terashima and T. Uesugi,
2211: Brane-antibrane action from boundary string field theory, JHEP 0103:
2212: 019 (2001) [arXiv:hep-th/0012210].
2213: 
2214: \bibitem{Polchinski}
2215: J.~Polchinski, ``M-theory and the light cone,'' Prog.\ Theor.\
2216: Phys.\ Suppl.\  {\bf 134}, 158 (1999) [arXiv:hep-th/9903165].
2217: 
2218: \bibitem{Guijosa} A.~Guijosa, H.~H.~Hernandez Hernandez and
2219: H.~A.~Morales Tecotl, ``The entropy of the rotating charged black
2220: threebrane from a brane-antibrane system,'' arXiv:hep-th/0402158.
2221: 
2222: \end{thebibliography}
2223: %====================================================================+
2224: \end{document}
2225: 
2226: