1:
2: \input harvmac.tex
3: \input epsf
4: %\draftmode
5:
6:
7: %-------------------------
8: % This paper uses harvmac
9: %-------------------------
10:
11: \font\ticp=cmcsc10
12: \font\cmss=cmss10 \font\cmsss=cmss10 at 7pt
13: \def\Title#1#2{\rightline{#1}\ifx\answ\bigans\nopagenumbers\pageno0\vskip1in
14: \else\pageno1\vskip.8in\fi \centerline{\titlefont #2}\vskip .5in}
15: \font\titlerm=cmr10 scaled\magstep3 \font\titlerms=cmr7 scaled\magstep3
16: \font\titlermss=cmr5 scaled\magstep3 \font\titlei=cmmi10 scaled\magstep3
17: \font\titleis=cmmi7 scaled\magstep3 \font\titleiss=cmmi5 scaled\magstep3
18: \font\titlesy=cmsy10 scaled\magstep3 \font\titlesys=cmsy7 scaled\magstep3
19: \font\titlesyss=cmsy5 scaled\magstep3 \font\titleit=cmti10 scaled\magstep3
20: \font\ticp=cmcsc10
21: \font\ttsmall=cmtt10 at 8pt
22:
23: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
24: %The following lines are needed to insert the accompanying figures
25: %in the paper. If you do not have epsf, then comment out the line
26: % ``\input epsf'', and print the figures separately.
27: \input epsf
28: %
29: \ifx\epsfbox\UnDeFiNeD\message{(NO epsf.tex, FIGURES WILL BE
30: IGNORED)}
31: \def\figin#1{\vskip2in}% blank space instead
32: \else\message{(FIGURES WILL BE INCLUDED)}\def\figin#1{#1}\fi
33: %
34: \def\ifig#1#2#3{\xdef#1{Fig.\the\figno}
35: \goodbreak\topinsert\figin{\centerline{#3}}%
36: \smallskip\centerline{\vbox{\baselineskip12pt
37: \advance\hsize by -1truein\noindent{\bf Fig.~\the\figno:} #2}}
38: \bigskip\endinsert\global\advance\figno by1}
39: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
40:
41:
42:
43:
44:
45: \baselineskip14pt
46: \def\T{$T^{1,1}$}
47: \def\no{ \vskip.15cm \noindent}
48: \def\O{ $\Omega$}
49: \def\N{\overline{N}}
50: \def\t{\times}
51: \def\tt{$\times$}
52: \def\p{\partial}
53: \def\pb{\bar{\partial}}
54: \def\zb{\bar{z}}
55: \def\Jt{\tilde{J}}
56: \def\ra{\rightarrow}
57: \def\ads{AdS$_3$}
58: \def\eps{$i\epsilon$}
59:
60: %
61: % definitions
62: %
63: \def\[{\left [}
64: \def\]{\right ]}
65: \def\({\left (}
66: \def\){\right )}
67: \def\itm{\noindent $\bullet$ \ }
68: \def\p{\partial}
69: \def\l{\ell}
70: \def\g{\gamma}
71: \def\d{\delta}
72: \def\e{\varepsilon}
73: \def\Om{\Omega}
74: \def\th{\theta}
75: \def\a{\alpha}
76: \def\lam{\lambda}
77: \def\CN{{\cal N}}
78: \def\CO{{\cal O}}
79: \def\dua{\( {\p \over \p u} \)^a}
80: \def\dva{\( {\p \over \p v} \)^a}
81: \def\dta{\( {\p \over \p t} \)^a}
82: \def\dt{\( {\p \over \p t} \)}
83: \def\dra{\( {\p \over \p r} \)^a}
84: \def\dza{\( {\p \over \p z} \)^a}
85: \def\dta{\( {\p \over \p t} \)^a}
86: \def\dpha{\( {\p \over \p \phi} \)^a}
87: \def\ud{\dot u}
88: \def\vd{\dot v}
89: \def\rd{\dot r}
90: \def\td{\dot t}
91: \def\phd{\dot {\phi}}
92: \def\sads{AdS Schwarzschild}
93: \def\rh{r_+}
94: \def\len{{\cal L}}
95: \def\lenr{{\cal L}_{\rm{reg}}}
96: \def\rinf{r_{\rm{max}}}
97: \def\ri{r_i}
98: \def\tc{t_c}
99: \def\Ec{E_c}
100: \def\rmi{r_-}
101: \def\rs{r_*}
102: \def\vv{\tilde{v}}
103: \def\uu{\tilde{u}}
104: \def\rmin{r_{\rm{min}}}
105: \def\half{{1 \over 2}}
106: \def\third{{1 \over 3}}
107: \def\quarter{{1 \over 4}}
108: \def\ft{{4 \over 3}}
109: \def\Rrr{(2 \rh^2 + R^2)}
110: \def\12{{1 \over 2}}
111: \def\Re{{\bf \rm Re}}
112: \def\Im{{\bf \rm Im}}
113: \def\AS5{{\rm Schw-AdS_5}}
114:
115: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
116:
117:
118: %--------------------------------------------------------------------------------
119:
120:
121: %\KloschQV
122: \lref\KloschQV{
123: T.~Klosch and T.~Strobl,
124: %``Classical and Quantum Gravity in 1+1 Dimensions, Part II: The Universal Coverings,''
125: Class.\ Quant.\ Grav.\ {\bf 13}, 2395 (1996)
126: [arXiv:gr-qc/9511081].
127: %%CITATION = GR-QC 9511081;%%
128: }
129:
130:
131: %\LeviCX
132: \lref\LeviCX{
133: T.~S.~Levi and S.~F.~Ross,
134: ``Holography beyond the horizon and cosmic censorship,''
135: arXiv:hep-th/0304150.
136: %%CITATION = HEP-TH 0304150;%%
137: }
138:
139:
140:
141:
142: \lref\Baker{ G.A. Baker and P.R. Graves-Morris, { \it Pad{\'e}
143: Approximants }, second edition, Cambridge Univ. Press, (1995).}
144:
145:
146: %\IchinoseRG
147: \lref\IchinoseRG{ I.~Ichinose and Y.~Satoh, ``Entropies of scalar
148: fields on three-dimensional black holes,'' Nucl.\ Phys.\ B {\bf
149: 447}, 340 (1995) [arXiv:hep-th/9412144].
150: %%CITATION = HEP-TH 9412144;%%
151: }
152:
153:
154:
155: %\ShenkerXQ
156: \lref\ShenkerXQ{ S.~H.~Shenker, ``Another Length Scale in String
157: Theory?,'' arXiv:hep-th/9509132.
158: %%CITATION = HEP-TH 9509132;%%
159: }
160:
161: %\KabatCU
162: \lref\KabatCU{ D.~Kabat and P.~Pouliot, ``A Comment on Zero-brane
163: Quantum Mechanics,'' Phys.\ Rev.\ Lett.\ {\bf 77}, 1004 (1996)
164: [arXiv:hep-th/9603127].
165: %%CITATION = HEP-TH 9603127;%%
166: }
167:
168:
169:
170: %\DouglasYP
171: \lref\DouglasYP{ M.~R.~Douglas, D.~Kabat, P.~Pouliot and
172: S.~H.~Shenker, ``D-branes and short distances in string theory,''
173: Nucl.\ Phys.\ B {\bf 485}, 85 (1997) [arXiv:hep-th/9608024].
174: %%CITATION = HEP-TH 9608024;%%
175: }
176:
177:
178:
179:
180: %\GaiottoRM
181: \lref\GaiottoRM{ D.~Gaiotto, N.~Itzhaki and L.~Rastelli, ``Closed
182: strings as imaginary D-branes,'' arXiv:hep-th/0304192.
183: %%CITATION = HEP-TH 0304192;%%
184: }
185:
186:
187:
188:
189: %\LambertZR
190: \lref\LambertZR{ N.~Lambert, H.~Liu and J.~Maldacena, `Closed
191: strings from decaying D-branes,'' arXiv:hep-th/0303139.
192: %%CITATION = HEP-TH 0303139;%%
193: }
194:
195:
196: %\MaldacenaBW
197: \lref\MaldacenaBW{ J.~M.~Maldacena and A.~Strominger, ``AdS(3)
198: black holes and a stringy exclusion principle,'' JHEP {\bf 9812},
199: 005 (1998) [arXiv:hep-th/9804085].
200: %%CITATION = HEP-TH 9804085;%%
201: }
202:
203:
204:
205: %\GubserNZ
206: \lref\GubserNZ{ S.~S.~Gubser, I.~R.~Klebanov and A.~A.~Tseytlin,
207: ``Coupling constant dependence in the thermodynamics of N = 4
208: supersymmetric Yang-Mills theory,'' Nucl.\ Phys.\ B {\bf 534}, 202
209: (1998) [arXiv:hep-th/9805156].
210: %%CITATION = HEP-TH 9805156;%%
211: }
212:
213:
214:
215:
216: %\GrisaruZN
217: \lref\GrisaruZN{ M.~T.~Grisaru, R.~C.~Myers and O.~Tafjord, ``SUSY
218: and Goliath,'' JHEP {\bf 0008}, 040 (2000) [arXiv:hep-th/0008015].
219: %%CITATION = HEP-TH 0008015;%%
220: }
221:
222:
223:
224:
225: %\McGreevyCW
226: \lref\McGreevyCW{ J.~McGreevy, L.~Susskind and N.~Toumbas,
227: ``Invasion of the giant gravitons from anti-de Sitter space,''
228: JHEP {\bf 0006}, 008 (2000) [arXiv:hep-th/0003075].
229: %%CITATION = HEP-TH 0003075;%%
230: }
231:
232:
233:
234:
235: %\NunezEQ
236: \lref\NunezEQ{ A.~Nunez and A.~O.~Starinets, ``AdS/CFT
237: correspondence, quasinormal modes, and thermal correlators in N =
238: 4 SYM,'' arXiv:hep-th/0302026.
239: %%CITATION = HEP-TH 0302026;%%
240: }
241:
242:
243: \lref\realtime{For a review see M. Le Bellac, {\it Thermal Field
244: Theory}, Cambridge University Press (1996).}
245: %\SusskindKW
246: \lref\SusskindKW{ L.~Susskind, ``The anthropic landscape of string
247: theory,'' arXiv:hep-th/0302219.
248: %%CITATION = HEP-TH 0302219;%%
249: }
250:
251: %\BerkoozJE
252: \lref\BerkoozJE{ M.~Berkooz, B.~Craps, D.~Kutasov and G.~Rajesh,
253: ``Comments on cosmological singularities in string theory,'' JHEP
254: {\bf 0303}, 031 (2003) [arXiv:hep-th/0212215].
255: %%CITATION = HEP-TH 0212215;%%
256: }
257:
258:
259: %\GiveonGE
260: \lref\GiveonGE{ A.~Giveon, E.~Rabinovici and A.~Sever, ``Beyond
261: the singularity of the 2-D charged black hole,''
262: arXiv:hep-th/0305140.
263: %%CITATION = HEP-TH 0305140;%%
264: }
265: %\GiveonGB
266: \lref\GiveonGB{ A.~Giveon, E.~Rabinovici and A.~Sever, ``Strings
267: in singular time-dependent backgrounds,'' arXiv:hep-th/0305137.
268: %%CITATION = HEP-TH 0305137;%%
269: }
270:
271:
272:
273: %\LoukoTP
274: \lref\LoukoTP{ J.~Louko, D.~Marolf and S.~F.~Ross, ``On geodesic
275: propagators and black hole holography,'' Phys.\ Rev.\ D {\bf 62},
276: 044041 (2000) [arXiv:hep-th/0002111].
277: %%CITATION = HEP-TH 0002111;%%
278: }
279:
280:
281: \lref\etinf{For reviews see A. Linde,
282: {\it Particle Physics and Inflationary Cosmology}
283: Harwood Academic, (1990), and A.~H.~Guth, ``Inflation and eternal
284: inflation,'' Phys.\ Rept.\ {\bf 333}, 555 (2000)
285: [arXiv:astro-ph/0002156]}
286:
287: \lref\BerryMI{ M. V. Berry,
288: ``Uniform asymptotic smoothing of
289: Stokes's discontinuities," Proc.\ Roy.\ Soc.\ Lond., {\bf A422}, 7
290: (1989)~.}
291:
292: \lref\BerryMII{ M. V. Berry and C. J. Howls,
293: ``Hyperasymptotics for integrals
294: with saddles," Proc.\ Roy.\ Soc.\ Lond., {\bf A434}, 657 (1991)~.}
295:
296: %\GrossGK
297: \lref\GrossGK{ D.~J.~Gross and H.~Ooguri, ``Aspects of large N
298: gauge theory dynamics as seen by string theory,'' Phys.\ Rev.\ D
299: {\bf 58}, 106002 (1998) [arXiv:hep-th/9805129].
300: %%CITATION = HEP-TH 9805129;%%
301: }
302:
303: %\StephensAN
304: \lref\StephensAN{ C.~R.~Stephens, G.~'t Hooft and B.~F.~Whiting,
305: ``Black hole evaporation without information loss,'' Class.\
306: Quant.\ Grav.\ {\bf 11}, 621 (1994) [arXiv:gr-qc/9310006].
307: %%CITATION = GR-QC 9310006;%%
308: }
309:
310: %\SusskindIF
311: \lref\SusskindIF{ L.~Susskind, L.~Thorlacius and J.~Uglum, ``The
312: Stretched horizon and black hole complementarity,'' Phys.\ Rev.\ D
313: {\bf 48}, 3743 (1993) [arXiv:hep-th/9306069].
314: %%CITATION = HEP-TH 9306069;%%
315: }
316:
317: %
318: %\KrausIV
319: \lref\KrausIV{ P.~Kraus, H.~Ooguri and S.~Shenker, ``Inside the
320: horizon with AdS/CFT,'' arXiv:hep-th/0212277.
321: %%CITATION = HEP-TH 0212277;%%
322: }
323:
324:
325:
326: %\InamiWU
327: \lref\Inami{ T.~Inami and H.~Ooguri, ``One Loop Effective
328: Potential In Anti-De Sitter Space,'' Prog.\ Theor.\ Phys.\ {\bf
329: 73}, 1051 (1985).
330: %%CITATION = PTPKA,73,1051;%%
331: }
332:
333: %\BurgessTI
334: \lref\Burgess{ C.~P.~Burgess and C.~A.~Lutken, ``Propagators And
335: Effective Potentials In Anti-De Sitter Space,'' Phys.\ Lett.\ B
336: {\bf 153}, 137 (1985).
337: %%CITATION = PHLTA,B153,137;%%
338: }
339:
340:
341:
342:
343: %\MaldacenaHW
344: \lref\MaldacenaHW{ J.~M.~Maldacena and H.~Ooguri, ``Strings in
345: AdS(3) and SL(2,R) WZW model. I,'' J.\ Math.\ Phys.\ {\bf 42},
346: 2929 (2001) [arXiv:hep-th/0001053]; Phys.\ Rev.\ D {\bf 65},
347: 106006 (2002) [arXiv:hep-th/0111180]; J.~M.~Maldacena, H.~Ooguri
348: and J.~Son, ``Strings in AdS(3) and the SL(2,R) WZW model. II:
349: Euclidean black
350: %hole,''
351: J.\ Math.\ Phys.\ {\bf 42}, 2961 (2001) [arXiv:hep-th/0005183];
352: %%CITATION = HEP-TH 0001053;%%
353: %\MaldacenaKM
354: %\lref\MaldacenaKM{
355: J.~M.~Maldacena and H.~Ooguri, ``Strings in AdS(3) and the SL(2,R)
356: WZW model. III: Correlation functions,'' Phys.\ Rev.\ D {\bf 65},
357: 106006 (2002) [arXiv:hep-th/0111180].
358: %%CITATION = HEP-TH 0111180;%%
359: }
360:
361:
362:
363:
364:
365: \lref\Ben{M. Berkooz, B. Craps, D. Kutasov, and G. Rajesh,
366: arXiv:hep-th/0212215.}
367:
368:
369:
370: %\Tolley
371: \lref\Tolley{ A.~J.~Tolley and N.~Turok, ``Quantum fields in a big
372: crunch / big bang spacetime,'' Phys.\ Rev.\ D {\bf 66}, 106005
373: (2002) [arXiv:hep-th/0204091].
374: %%CITATION = HEP-TH 0204091;%%
375: }
376:
377: %\Herzog
378: \lref\Herzog{ C.~P.~Herzog and D.~T.~Son, ``Schwinger-Keldysh
379: Propagators from AdS/CFT Correspondence,'' arXiv:hep-th/0212072.
380: %%CITATION = HEP-TH 0212072;%%
381: }
382:
383:
384:
385:
386: %%\MaldacenaKV
387: %\lref\MaldacenaKV{
388: %J.~M.~Maldacena, H.~Ooguri and J.~Son,
389: %``Strings in AdS(3) and the SL(2,R) WZW model. II: Euclidean black
390: %hole,''
391: %J.\ Math.\ Phys.\ {\bf 42}, 2961 (2001)
392: %[arXiv:hep-th/0005183].
393: %%%CITATION = HEP-TH 0005183;%%
394: %}
395:
396: %\KutasovXU
397: \lref\KutasovXU{ D.~Kutasov and N.~Seiberg, ``More comments on
398: string theory on AdS(3),'' JHEP {\bf 9904}, 008 (1999)
399: [arXiv:hep-th/9903219].
400: %%CITATION = HEP-TH 9903219;%%
401: }
402:
403: %\deBoerPP
404: \lref\deBoerPP{ J.~de Boer, H.~Ooguri, H.~Robins and
405: J.~Tannenhauser, ``String theory on AdS(3),'' JHEP {\bf 9812}, 026
406: (1998) [arXiv:hep-th/9812046].
407: %%CIT
408: }
409:
410: %%\MaldacenaKM
411: %\lref\MaldacenaKM{
412: %J.~M.~Maldacena and H.~Ooguri,
413: %``Strings in AdS(3) and the SL(2,R) WZW model. III: Correlation
414: %functions,''
415: %Phys.\ Rev.\ D {\bf 65}, 106006 (2002)
416: %[arXiv:hep-th/0111180].
417: %%%CITATION = HEP-TH 0111180;%%
418: %}
419:
420: %\CruzIR
421: \lref\CruzIR{ N.~Cruz, C.~Martinez and L.~Pena, ``Geodesic
422: Structure Of The (2+1) Black Hole,''
423: %Class.\ Quant.\ Grav.\ {\bf 11}, 2731 (1994)
424: [arXiv:gr-qc/9401025].
425: %%CITATION = GR-QC 9401025;%%
426: }
427:
428: %\MartinecXQ
429: \lref\MartinecXQ{ E.~J.~Martinec and W.~McElgin, ``Exciting AdS
430: orbifolds,'' arXiv:hep-th/0206175.
431: %%CITATION = HEP-TH 0206175;%%
432: }
433:
434: %\GiveonNS
435: \lref\GiveonNS{ A.~Giveon, D.~Kutasov and N.~Seiberg, ``Comments
436: on string theory on AdS(3),'' Adv.\ Theor.\ Math.\ Phys.\ {\bf
437: 2}, 733 (1998) [arXiv:hep-th/9806194].
438: %%CITATION = HEP-TH 9806194;%%
439: }
440:
441: %\SonQM
442: \lref\SonQM{ J.~Son, ``String theory on AdS(3)/Z(N),''
443: arXiv:hep-th/0107131.
444: %%CITATION = HEP-TH 0107131;%%
445: }
446:
447: %\HananyEV
448: \lref\HananyEV{ A.~Hanany, N.~Prezas and J.~Troost, ``The
449: partition function of the two-dimensional black hole conformal
450: %field theory,''
451: JHEP {\bf 0204}, 014 (2002) [arXiv:hep-th/0202129].
452: %%CITATION = HEP-TH 0202129;%%
453: }
454:
455: %\TeschnerFT
456: \lref\TeschnerFT{ J.~Teschner, Nucl.\ Phys.\ B {\bf 546}, 390
457: (1999) [arXiv:hep-th/9712256];
458: %%CITATION = HEP-TH 9712256;%%
459: %
460: %\TeschnerUG
461: %\lref\TeschnerUG{
462: %J.~Teschner,
463: Nucl.\ Phys.\ B {\bf 571}, 555 (2000) [arXiv:hep-th/9906215].
464: %%CITATION = HEP-TH 9906215;%%
465: }
466:
467:
468: %\DanielssonZT
469: \lref\DanielssonZT{ U.~H.~Danielsson, E.~Keski-Vakkuri and
470: M.~Kruczenski, ``Spherically collapsing matter in AdS, holography,
471: and shellons,'' Nucl.\ Phys.\ B {\bf 563}, 279 (1999)
472: [arXiv:hep-th/9905227].
473: %%CITATION = HEP-TH 9905227;%%
474: }
475: %\EvansFR
476: \lref\EvansFR{ T.~S.~Evans, A.~Gomez Nicola, R.~J.~Rivers and
477: D.~A.~Steer, ``Transport coefficients and analytic continuation in
478: dual 1+1
479: %dimensional models at finite temperature,''
480: arXiv:hep-th/0204166.
481: %%CITATION = HEP-TH 0204166;%%
482: }
483:
484:
485: %\BanadosWN
486: \lref\BanadosWN{ M.~Banados, C.~Teitelboim and J.~Zanelli, ``The
487: Black Hole In Three-Dimensional Space-Time,'' Phys.\ Rev.\ Lett.\
488: {\bf 69}, 1849 (1992) [arXiv:hep-th/9204099];
489: %%CITATION = HEP-TH 9204099;%%
490: %\BanadosGQ
491: %\lref\BanadosGQ{
492: M.~Banados, M.~Henneaux, C.~Teitelboim and J.~Zanelli, ``Geometry
493: of the (2+1) black hole,'' Phys.\ Rev.\ D {\bf 48}, 1506 (1993)
494: [arXiv:gr-qc/9302012].}
495:
496: %\IsraelUR
497: \lref\IsraelUR{ W.~Israel, ``Thermo Field Dynamics Of Black
498: Holes,'' Phys.\ Lett.\ A {\bf 57}, 107 (1976).
499: %%CITATION = PHLTA,A57,107;%%
500: }
501:
502: %\UnruhDB
503: \lref\UnruhDB{ W.~G.~Unruh, ``Notes On Black Hole Evaporation,''
504: Phys.\ Rev.\ D {\bf 14}, 870 (1976).
505: %%CITATION = PHRVA,D14,870;%%
506: }
507:
508:
509:
510: %\ElitzurRT
511: \lref\ElitzurRT{ S.~Elitzur, A.~Giveon, D.~Kutasov and
512: E.~Rabinovici, ``From big bang to big crunch and beyond,'' JHEP
513: {\bf 0206}, 017 (2002) [arXiv:hep-th/0204189].
514: %%CITATION = HEP-TH 0204189;%%
515: }
516:
517: %\CrapsII
518: \lref\CrapsII{ B.~Craps, D.~Kutasov and G.~Rajesh, ``String
519: propagation in the presence of cosmological singularities,'' JHEP
520: {\bf 0206}, 053 (2002) [arXiv:hep-th/0205101].
521: %%CITATION = HEP-TH 0205101;%%
522: }
523:
524: %\CornalbaNV
525: \lref\CornalbaNV{ L.~Cornalba, M.~S.~Costa and C.~Kounnas, ``A
526: resolution of the cosmological singularity with orientifolds,''
527: arXiv:hep-th/0204261.
528: %%CITATION = HEP-TH 0204261;%%
529: }
530:
531: %\GubserBC
532: \lref\GubserBC{ S.~S.~Gubser, I.~R.~Klebanov and A.~M.~Polyakov,
533: ``Gauge theory correlators from non-critical string theory,''
534: Phys.\ Lett.\ B {\bf 428}, 105 (1998) [arXiv:hep-th/9802109].
535: %%CITATION = HEP-TH 9802109;%%
536: }
537:
538: %\KeskiVakkuriNW
539: \lref\KeskiVakkuriNW{ E.~Keski-Vakkuri, ``Bulk and boundary
540: dynamics in BTZ black holes,'' Phys.\ Rev.\ D {\bf 59}, 104001
541: (1999) [arXiv:hep-th/9808037].
542: %%CITATION = HEP-TH 9808037;%%
543: }
544:
545:
546:
547: %\MaldacenaRE
548: \lref\MaldacenaRE{ J.~M.~Maldacena, ``The large N limit of
549: superconformal field theories and supergravity,'' Adv.\ Theor.\
550: Math.\ Phys.\ {\bf 2}, 231 (1998) [Int.\ J.\ Theor.\ Phys.\ {\bf
551: 38}, 1113 (1999)] [arXiv:hep-th/9711200].
552: %%CITATION = HEP-TH 9711200;%%
553: }
554:
555:
556: %\LifschytzEB
557: \lref\LifschytzEB{ G.~Lifschytz and M.~Ortiz, ``Scalar Field
558: Quantization On The (2+1)-Dimensional Black Hole
559: %Background,''
560: Phys.\ Rev.\ D {\bf 49}, 1929 (1994) [arXiv:gr-qc/9310008].
561: %%CITATION
562: }
563:
564: %\BalasubramanianSN
565: \lref\BalasubramanianSN{ V.~Balasubramanian, P.~Kraus and
566: A.~E.~Lawrence, ``Bulk vs. boundary dynamics in anti-de Sitter
567: spacetime,'' Phys.\ Rev.\ D {\bf 59}, 046003 (1999)
568: [arXiv:hep-th/9805171].
569: %%CITATION = HEP-TH 9805171;%%
570: }
571:
572: %\WittenQJ
573: \lref\WittenQJ{ E.~Witten, ``Anti-de Sitter space and
574: holography,'' Adv.\ Theor.\ Math.\ Phys.\ {\bf 2}, 253 (1998)
575: [arXiv:hep-th/9802150].
576: %%CITATION = HEP-TH 9802150;%%
577: }
578:
579: %\BalasubramanianRE
580: \lref\BalasubramanianRE{ V.~Balasubramanian and P.~Kraus, ``A
581: stress tensor for anti-de Sitter gravity,'' Commun.\ Math.\ Phys.\
582: {\bf 208}, 413 (1999) [arXiv:hep-th/9902121].
583: %%CITATION = HEP-TH 9902121;%%
584: }
585:
586:
587: %\HorowitzXK
588: \lref\HorowitzXK{ G.~T.~Horowitz and D.~Marolf, ``A new approach
589: to string cosmology,'' JHEP {\bf 9807}, 014 (1998)
590: [arXiv:hep-th/9805207].
591: %%CITATION = HEP-TH 9805207;%%
592: }
593:
594: %\BalasubramanianDE
595: \lref\BalasubramanianDE{ V.~Balasubramanian, P.~Kraus,
596: A.~E.~Lawrence and S.~P.~Trivedi, ``Holographic probes of anti-de
597: Sitter space-times,'' Phys.\ Rev.\ D {\bf 59}, 104021 (1999)
598: [arXiv:hep-th/9808017].
599: %%CITATION = HEP-TH 9808017;%%
600: }
601:
602: %\CarneirodaCunhaNW
603: \lref\CarneirodaCunhaNW{ B.~G.~Carneiro da Cunha,
604: ``Three-dimensional de Sitter gravity and the correspondence,''
605: Phys.\ Rev.\ D {\bf 65}, 104025 (2002) [arXiv:hep-th/0110169].
606: %%CITATION = HEP-TH 0110169;%%
607: }
608:
609:
610: %\MaldacenaKR
611: \lref\MaldacenaKR{ J.~M.~Maldacena, ``Eternal black holes in
612: Anti-de-Sitter,'' arXiv:hep-th/0106112.
613: %%CITATION = HEP-TH 0106112;%%
614: }
615:
616:
617: %\GiveonNS
618: \lref\GiveonNS{ A.~Giveon, D.~Kutasov and N.~Seiberg, ``Comments
619: on string theory on AdS(3),'' Adv.\ Theor.\ Math.\ Phys.\ {\bf
620: 2}, 733 (1998) [arXiv:hep-th/9806194].
621: %%CITATION
622: }
623:
624:
625: %\TeschnerFT
626: \lref\TeschnerFT{ J.~Teschner, ``On structure constants and fusion
627: rules in the SL(2,C)/SU(2) WZNW
628: %model,''
629: Nucl.\ Phys.\ B {\bf 546}, 390 (1999) [arXiv:hep-th/9712256];
630: ``Operator product expansion and factorization in the H-3+ WZNW
631: %model,''
632: Nucl.\ Phys.\ B {\bf 571}, 555 (2000) [arXiv:hep-th/9906215].
633: %%CITATION = HEP-TH 9906215;%%
634: }
635:
636: %\HorowitzJC
637: \lref\HorowitzJC{ G.~T.~Horowitz and D.~L.~Welch, ``Exact
638: three-dimensional black holes in string theory,'' Phys.\ Rev.\
639: Lett.\ {\bf 71}, 328 (1993) [arXiv:hep-th/9302126];
640: %%%CITATION = HEP-TH 9302126;%%
641: %}
642: %
643: %}
644: %%\KaloperKJ
645: %\lref\KaloperKJ{
646: N.~Kaloper,
647: %``Miens of the three-dimensional black hole,''
648: Phys.\ Rev.\ D {\bf 48}, 2598 (1993) [arXiv:hep-th/9303007];
649: %%%CITATION = HEP-TH 9303007;%%
650: %}
651: %
652: %\NatsuumeIJ
653: %\lref\NatsuumeIJ{
654: M.~Natsuume and Y.~Satoh, ``String theory on three dimensional
655: black holes,'' Int.\ J.\ Mod.\ Phys.\ A {\bf 13}, 1229 (1998)
656: [arXiv:hep-th/9611041];
657: %%%CITATION = HEP-TH 9611041;%%
658: %}
659: %\SatohXE
660: Y.~Satoh, ``Ghost-free and modular invariant spectra of a string
661: in SL(2,R) and
662: %three-dimensional black hole geometry,''
663: Nucl.\ Phys.\ B {\bf 513}, 213 (1998) [arXiv:hep-th/9705208];
664: %%CITATION = HEP-TH 9705208;%%
665: %}
666: %
667: %%\MaldacenaBW
668: %\lref\MaldacenaBW{
669: J.~M.~Maldacena and A.~Strominger, ``AdS(3) black holes and a
670: stringy exclusion principle,'' JHEP {\bf 9812}, 005 (1998)
671: [arXiv:hep-th/9804085];
672: %%%CITATION = HEP-TH 9804085;%%
673: %}
674: %
675: %%\HemmingWE
676: %\lref\HemmingWE{
677: S.~Hemming and E.~Keski-Vakkuri,
678: %``The spectrum of strings on BTZ black holes and spectral flow in
679: %the
680: %SL(2,R) WZW model,''
681: Nucl.\ Phys.\ B {\bf 626}, 363 (2002) [arXiv:hep-th/0110252];
682: %%%CITATION = HEP-TH 0110252;%%
683: %}
684: %
685: %%\TroostWK
686: %\lref\TroostWK{
687: J.~Troost,
688: ``Winding strings and AdS(3) black holes,'' arXiv:hep-th/0206118;
689: %%CITATION = HEP-TH 0206118;%%
690: %}
691: %
692: %%\MartinecCF
693: %\lref\MartinecCF{
694: E.~J.~Martinec and W.~McElgin, ``String theory on AdS orbifolds,''
695: JHEP {\bf 0204}, 029 (2002) [arXiv:hep-th/0106171];
696: %%%CITATION = HEP-TH 0106171;%%
697: }
698:
699:
700: %\BalasubramanianRY
701: \lref\BalasubramanianRY{ V.~Balasubramanian, S.~F.~Hassan,
702: E.~Keski-Vakkuri and A.~Naqvi, ``A space-time orbifold: A toy
703: model for a cosmological
704: %singularity,''
705: arXiv:hep-th/0202187.
706: %%CITATION = HEP-TH 0202187;%%
707: }
708:
709: %\CornalbaFI
710: \lref\CornalbaFI{ L.~Cornalba and M.~S.~Costa, ``A New
711: Cosmological Scenario in String Theory,'' arXiv:hep-th/0203031.
712: %%CITATION = HEP-TH 0203031;%%
713: }
714:
715: %\NekrasovKF
716: \lref\NekrasovKF{ N.~Nekrasov, ``Milne universe, tachyons, and
717: quantum group'' arXiv: hep-th/0203112.
718: %%CITATION = HEP-TH 0203112;%%
719: }
720:
721: %\SimonMA
722: \lref\SimonMA{ J.~Simon, ``The geometry of null rotation
723: identifications,'' JHEP {\bf 0206}, 001 (2002)
724: [arXiv:hep-th/0203201].
725: %%CITATI
726: }
727:
728: %\LiuFT
729: \lref\LiuFT{ H.~Liu, G.~Moore and N.~Seiberg, ``Strings in a
730: time-dependent orbifold,'' JHEP {\bf 0206}, 045 (2002)
731: [arXiv:hep-th/0204168];
732: %%CITATION = HEP-TH 0204168;%%
733: %\LiuKB
734: %\lref\LiuKB{
735: %H.~Liu, G.~Moore and N.~Seiberg,
736: ``Strings in time-dependent orbifolds,'' arXiv:hep-th/0206182.
737: %%CITATION = HEP-TH 02061
738: }
739:
740: %\LawrenceAJ
741: \lref\LawrenceAJ{ A.~Lawrence, ``On the instability of 3D null
742: singularities,'' arXiv:hep-th/0205288.
743: %%CITATION = HEP-TH 0205288;%%
744: }
745:
746: %\FabingerKR
747: \lref\FabingerKR{ M.~Fabinger and J.~McGreevy, ``On smooth
748: time-dependent orbifolds and null singularities,''
749: arXiv:hep-th/0206196.
750: %%CITATION = HEP-TH 0206196;%%
751: }
752:
753: %\HorowitzMW
754: \lref\HorowitzMW{ G.~T.~Horowitz and J.~Polchinski, ``Instability
755: of spacelike and null orbifold singularities,''
756: arXiv:hep-th/0206228.
757: %%CITATION = HEP-TH 0206228;%%
758: }
759:
760: %\SusskindQC
761: \lref\SusskindQC{ L.~Susskind and J.~Uglum, ``String Physics and
762: Black Holes,'' Nucl.\ Phys.\ Proc.\ Suppl.\ {\bf 45BC}, 115
763: (1996) [arXiv:hep-th/9511227].
764: %%CITATION = HEP-TH 9511227;%%
765: }
766:
767: %\HartleAI
768: \lref\HartleAI{ J.~B.~Hartle and S.~W.~Hawking, ``Wave Function Of
769: The Universe,'' Phys.\ Rev.\ D {\bf 28}, 2960 (1983).
770: %%CITATION = PHRVA,D28,2960;%%
771: }
772:
773: %\HartleTP
774: \lref\HartleTP{ J.~B.~Hartle and S.~W.~Hawking, ``Path Integral
775: Derivation Of Black Hole Radiance,'' Phys.\ Rev.\ D {\bf 13}, 2188
776: (1976).
777: %%CITATION = PHRVA,D13,2188;%%
778: }
779:
780: %\NiemiNF
781: \lref\NiemiNF{ A.~J.~Niemi and G.~W.~Semenoff, ``Finite
782: Temperature Quantum Field Theory In Minkowski Space,'' Annals
783: Phys.\ {\bf 152}, 105 (1984).
784: %%CITATION = APNYA,152,105;%%
785: }
786:
787: %\HemmingKD
788: \lref\HemmingKD{ S.~Hemming, E.~Keski-Vakkuri and P.~Kraus,
789: ``Strings in the extended BTZ spacetime,'' JHEP {\bf 0210}, 006
790: (2002) [arXiv:hep-th/0208003].
791: %%CITATION = HEP-TH 0208003;%%
792: }
793:
794: %\DixonJW
795: \lref\DixonJW{ L.~J.~Dixon, J.~A.~Harvey, C.~Vafa and E.~Witten,
796: ``Strings On Orbifolds,'' Nucl.\ Phys.\ B {\bf 261}, 678 (1985).
797: %%CITATION = NUPHA,B261,678;%%
798: }
799:
800: %\StromingerCZ
801: \lref\StromingerCZ{ A.~Strominger, ``Massless black holes and
802: conifolds in string theory,'' Nucl.\ Phys.\ B {\bf 451}, 96 (1995)
803: [arXiv:hep-th/9504090].
804: %%CITATION = HEP-TH 9504090;%%
805: }
806:
807: %\JohnsonQT
808: \lref\JohnsonQT{ C.~V.~Johnson, A.~W.~Peet and J.~Polchinski,
809: ``Gauge theory and the excision of repulson singularities,''
810: Phys.\ Rev.\ D {\bf 61}, 086001 (2000) [arXiv:hep-th/9911161].
811: %%CITATION = HEP-TH 9911161;%%
812: }
813:
814: %\BanksVH
815: \lref\BanksVH{ T.~Banks, W.~Fischler, S.~H.~Shenker and
816: L.~Susskind, ``M theory as a matrix model: A conjecture,'' Phys.\
817: Rev.\ D {\bf 55}, 5112 (1997) [arXiv:hep-th/9610043].
818: %%CITATION = HEP-TH 9610043;%%
819: }
820:
821: %\RastelliUV
822: \lref\RastelliUV{ For a review of some recent developments see
823: L.~Rastelli, A.~Sen and B.~Zwiebach, ``Vacuum string field
824: theory,'' arXiv:hep-th/0106010.
825: %%CITATION = HEP-TH 0106010;%%
826: }
827:
828: %\WittenZW
829: \lref\WittenZW{ E.~Witten, ``Anti-de Sitter space, thermal phase
830: transition, and confinement in gauge theories,'' Adv.\ Theor.\
831: Math.\ Phys.\ {\bf 2}, 505 (1998) [arXiv:hep-th/9803131].
832: %%CITATION = HEP-TH 9803131;%%
833: }
834:
835: %\BalasubramanianZV
836: \lref\BalasubramanianZV{ V.~Balasubramanian and S.~F.~Ross,
837: ``Holographic particle detection,'' Phys.\ Rev.\ D {\bf 61},
838: 044007 (2000) [arXiv:hep-th/9906226].
839: %%CITATION = HEP-TH 9906226;%%
840: }
841: %\HubenyDG
842: \lref\HubenyDG{ V.~E.~Hubeny, ``Precursors see inside black
843: holes,'' arXiv:hep-th/0208047.
844: %%CITATION = HEP-TH 0208047;%%
845: }
846:
847: %\SusskindIF
848: \lref\SusskindIF{ L.~Susskind, L.~Thorlacius and J.~Uglum, ``The
849: Stretched horizon and black hole complementarity,'' Phys.\ Rev.\ D
850: {\bf 48}, 3743 (1993) [arXiv:hep-th/9306069].
851: %%CITATION = HEP-TH 9306069;%%
852: }
853:
854: \lref\gks{N. Goheer, M. Kleban, and L. Susskind, ``The Trouble with de
855: Sitter Space,"
856: [arXiv:hep-th/0212209].}
857:
858: \lref\dks{L. Dyson, M. Kleban, and L. Susskind, ``Disturbing Implications of a Cosmological Constant,"
859: JHEP 0210:011,2002 [arXiv:hep-th/0208013].
860:
861: %\MaldacenaBW
862: \lref\MaldacenaBW{ J.~M.~Maldacena and A.~Strominger, ``AdS(3)
863: black holes and a stringy exclusion principle,'' JHEP {\bf 9812},
864: 005 (1998) [arXiv:hep-th/9804085].
865: %%CITATION = HEP-TH 9804085;%%
866: }
867:
868: %\DysonNT
869: \lref\DysonNT{ L.~Dyson, J.~Lindesay and L.~Susskind, ``Is there
870: really a de Sitter/CFT duality,'' JHEP {\bf 0208}, 045 (2002)
871: [arXiv:hep-th/0202163].}
872: %%CITATION = HEP-TH 0202163;%%
873: }
874:
875: %\WittenQJ
876: \lref\WittenQJ{
877: E.~Witten,
878: ``Anti-de Sitter space and holography,''
879: Adv.\ Theor.\ Math.\ Phys.\ {\bf 2}, 253 (1998)
880: [arXiv:hep-th/9802150].
881: %%CITATION = HEP-TH 9802150;%%
882: }
883:
884: %\AharonyTI
885: \lref\AharonyTI{
886: O.~Aharony, S.~S.~Gubser, J.~M.~Maldacena, H.~Ooguri and Y.~Oz,
887: ``Large N field theories, string theory and gravity,''
888: Phys.\ Rept.\ {\bf 323}, 183 (2000)
889: [arXiv:hep-th/9905111].
890: %%CITATION = HEP-TH 9905111;%%
891: }
892: %\GiddingsPT
893: \lref\GiddingsPT{
894: S.~B.~Giddings and M.~Lippert,
895: ``Precursors, black holes, and a locality bound,''
896: Phys.\ Rev.\ D {\bf 65}, 024006 (2002)
897: [arXiv:hep-th/0103231].
898: %%CITATION = HEP-TH 0103231;%%
899: }
900:
901: %\FreivogelEX
902: \lref\FreivogelEX{
903: B.~Freivogel, S.~B.~Giddings and M.~Lippert,
904: ``Toward a theory of precursors,''
905: Phys.\ Rev.\ D {\bf 66}, 106002 (2002)
906: [arXiv:hep-th/0207083].
907: %%CITATION = HEP-TH 0207083;%%
908: }
909:
910: %\HorowitzFM
911: \lref\HorowitzFM{
912: G.~T.~Horowitz and V.~E.~Hubeny,
913: ``CFT description of small objects in AdS,''
914: JHEP {\bf 0010}, 027 (2000)
915: [arXiv:hep-th/0009051].
916: %%CITATION = HEP-TH 0009051;%%
917: }
918:
919: %\JacobsonMI
920: \lref\JacobsonMI{
921: T.~Jacobson,
922: ``On the nature of black hole entropy,''
923: arXiv:gr-qc/9908031.
924: %%CITATION = GR-QC 9908031;%%
925: }
926:
927: %\BanksDD
928: \lref\BanksDD{
929: T.~Banks, M.~R.~Douglas, G.~T.~Horowitz and E.~J.~Martinec,
930: ``AdS dynamics from conformal field theory,''
931: arXiv:hep-th/9808016.
932: %%CITATION = HEP-TH 9808016;%%
933: }
934:
935: %\BalasubramanianDE
936: \lref\BalasubramanianDE{
937: V.~Balasubramanian, P.~Kraus, A.~E.~Lawrence and S.~P.~Trivedi,
938: ``Holographic probes of anti-de Sitter space-times,''
939: Phys.\ Rev.\ D {\bf 59}, 104021 (1999)
940: [arXiv:hep-th/9808017].
941: %%CITATION = HEP-TH 9808017;%%
942: }
943:
944: %\KabatYQ
945: \lref\KabatYQ{
946: D.~Kabat and G.~Lifschytz,
947: ``Gauge theory origins of supergravity causal structure,''
948: JHEP {\bf 9905}, 005 (1999)
949: [arXiv:hep-th/9902073].
950: %%CITATION = HEP-TH 9902073;%%
951: }
952:
953: %\DanielssonFA
954: \lref\DanielssonFA{
955: U.~H.~Danielsson, E.~Keski-Vakkuri and M.~Kruczenski,
956: ``Black hole formation in AdS and thermalization on the boundary,''
957: JHEP {\bf 0002}, 039 (2000)
958: [arXiv:hep-th/9912209].
959: %%CITATION = HEP-TH 9912209;%%
960: }
961:
962: %\GregoryAN
963: \lref\GregoryAN{
964: J.~P.~Gregory and S.~F.~Ross,
965: ``Looking for event horizons using UV/IR relations,''
966: Phys.\ Rev.\ D {\bf 63}, 104023 (2001)
967: [arXiv:hep-th/0012135].
968: %%CITATION = HEP-TH 0012135;%%
969: }
970:
971: %\SusskindEY
972: \lref\SusskindEY{
973: L.~Susskind and N.~Toumbas,
974: ``Wilson loops as precursors,''
975: Phys.\ Rev.\ D {\bf 61}, 044001 (2000)
976: [arXiv:hep-th/9909013].
977: %%CITATION = HEP-TH 9909013;%%
978: }
979:
980: %\HashimotoZP
981: \lref\HashimotoZP{
982: A.~Hashimoto, S.~Hirano and N.~Itzhaki,
983: ``Large branes in AdS and their field theory dual,''
984: JHEP {\bf 0008}, 051 (2000)
985: [arXiv:hep-th/0008016].
986: %%CITATION = HEP-TH 0008016;%%
987: }
988:
989: %\HawkingDH
990: \lref\HawkingDH{
991: S.~W.~Hawking and D.~N.~Page,
992: ``Thermodynamics Of Black Holes In Anti-De Sitter Space,''
993: Commun.\ Math.\ Phys.\ {\bf 87}, 577 (1983).
994: %%CITATION = CMPHA,87,577;%%
995: }
996:
997: %\SusskindVU
998: \lref\SusskindVU{ L.~Susskind, ``The World as a hologram,'' J.\
999: Math.\ Phys.\ {\bf 36}, 6377 (1995) [arXiv:hep-th/9409089].
1000: %%CITATION = HEP-TH 9409089;%%
1001: }
1002:
1003: %\GubserBC
1004: \lref\GubserBC{ S.~S.~Gubser, I.~R.~Klebanov and A.~M.~Polyakov,
1005: ``Gauge theory correlators from non-critical string theory,''
1006: Phys.\ Lett.\ B {\bf 428}, 105 (1998) [arXiv:hep-th/9802109].
1007: %%CITATION = HEP-TH 9802109;%%
1008: }
1009:
1010:
1011:
1012: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1013: \Title{\vbox{\baselineskip12pt \hbox{hep-th/0306170}
1014: \hbox{SU-ITP-03-16} \vskip-.4in}} {The Black Hole Singularity in
1015: AdS/CFT }
1016:
1017:
1018:
1019: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1020: \centerline{Lukasz Fidkowski, Veronika Hubeny, Matthew Kleban, and
1021: Stephen Shenker}
1022: \bigskip
1023:
1024: \centerline{\it Department of Physics, Stanford University,
1025: Stanford, CA, 94305, USA}
1026:
1027: \bigskip\bigskip \baselineskip14pt
1028:
1029:
1030:
1031: \noindent
1032:
1033: We explore physics behind the horizon in eternal AdS Schwarzschild
1034: black holes. In dimension $d >3$ ,
1035: %(where the singularity is locally Schwarzschild)
1036: %(where the curvature invariants blow up as one approaches the singularity),
1037: where the curvature grows large near the singularity, we find
1038: distinct but subtle signals of this singularity in the boundary
1039: CFT correlators. Building on previous work, we study correlation
1040: functions of operators on the two disjoint asymptotic boundaries
1041: of the spacetime by investigating the spacelike geodesics that
1042: join the boundaries. These dominate the correlators for large mass
1043: bulk fields. We show that the Penrose diagram for $d>3$
1044: is not square. As a result, the real geodesic connecting the two boundary points becomes
1045: almost null and bounces off the singularity at a finite boundary time $t_c \neq
1046: 0$.
1047:
1048: If this geodesic were to dominate the correlator there would be a
1049: ``light cone" singularity at $t_c$. However, general properties of the
1050: boundary theory rule this out. In fact, we argue that the
1051: correlator is actually dominated by a complexified geodesic, whose properties yield
1052: the large mass quasinormal mode frequencies previously
1053: found for this black hole.
1054: % Three geodesics coincide at $t=0$, giving rise to a branch cut in the
1055: % correlator at small time (in the limit of large mass).
1056: % OR: ---------------
1057: We find a branch cut in the correlator at small time
1058: (in the limit of large mass), arising from coincidence
1059: of three geodesics.
1060: %----------------
1061: The $t_c$ singularity, a signal of the black hole singularity, occurs on a
1062: secondary sheet of the analytically continued correlator. Its properties are
1063: computationally accessible.
1064:
1065: The $t_c$ singularity persists to all orders in the $1/m$
1066: expansion, for finite $\alpha'$, and to all orders in $g_s$.
1067: Certain leading nonperturbative effects can also be studied. The
1068: behavior of these boundary theory quantities near $t_c$ gives, in
1069: principle, significant information about stringy and quantum
1070: behavior in the vicinity of the black hole singularity.
1071:
1072: \Date{June, 2003}
1073: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1074:
1075:
1076:
1077:
1078:
1079: \newsec{Introduction}
1080:
1081: Some of the deepest mysteries in quantum gravity lie hidden behind
1082: horizons, including the nature of the spacelike singularities
1083: inside black holes. A better understanding of these may shed more
1084: light on cosmological singularities as well. Despite much recent
1085: work \refs{\BalasubramanianRY,\CornalbaFI,\NekrasovKF,\SimonMA,
1086: \LiuFT,\ElitzurRT,\CornalbaNV,\CrapsII,\FabingerKR,
1087: \LawrenceAJ,\HorowitzMW, \BerkoozJE, \GiveonGE, \GiveonGB }, these
1088: issues remain mysterious. More generally, the existence of
1089: horizons and resolving the puzzles of information loss and quantum
1090: decoherence require that the description of physics inside the
1091: horizon be linked somehow to the description outside. This
1092: connection--the central concept of black hole complementarity
1093: \refs{\StephensAN, \SusskindIF}--is still not well understood.
1094:
1095: As observations continue to bolster the case for cosmological
1096: inflation, this set of questions becomes more pressing. A
1097: complete description of eternal inflation \etinf\ probably
1098: requires an understanding of the many ``universes" behind the de
1099: Sitter horizon of a single inflating patch \refs{\SusskindKW, \dks, \gks }.
1100:
1101: Despite the enormous progress in our understanding of quantum
1102: gravity in recent years, physics behind the horizon remains
1103: extremely puzzling. This is due in part to the holographic
1104: \refs{\StephensAN , \SusskindVU} nature of our most fully
1105: developed nonperturbative descriptions of quantum gravity, the
1106: brane bulk correspondences of Matrix Theory \BanksVH\ and AdS/CFT
1107: \MaldacenaRE, \refs{\GubserBC ,\WittenQJ ,\AharonyTI}. These
1108: formulations describe physics from the point of view of an
1109: observer outside the horizon. For example, CFT correlators in
1110: AdS/CFT describe observables on the asymptotic spatial boundary
1111: of AdS space. In the thermal CFT that describes a large black hole
1112: in AdS/CFT \WittenZW, these asymptotic observables apparently
1113: describe physics only outside the horizon.
1114:
1115: The brane bulk correspondences provide a fully self consistent
1116: quantum description of gravitational dynamics outside of the
1117: horizon, demonstrating that quantum coherence is not lost. But
1118: these descriptions appear to have little to say about behind the
1119: horizon phenomena.
1120:
1121: However, this is not the whole story. A number of authors have
1122: explored ways to extract behind the horizon information from
1123: boundary correlators \refs{\BalasubramanianZV, \LoukoTP,
1124: \MaldacenaKR, \KrausIV, \LeviCX}.\foot{ There is also large body of work
1125: addressing more general questions of extracting behind the horizon
1126: physics and holographic representation of horizons, as well as of
1127: generic local objects in the bulk; see e.g.\
1128: \refs{\BanksDD,\BalasubramanianDE,\KabatYQ,\JacobsonMI,\SusskindEY,
1129: \DanielssonFA,\HorowitzFM,\GregoryAN,\GiddingsPT,\FreivogelEX,\HubenyDG}.}
1130: In particular Maldacena \MaldacenaKR\ argued that the boundary
1131: description of the eternal Anti-de Sitter Schwarzschild black hole
1132: consisting of one copy of the CFT on each of the two asymptotic
1133: boundaries could give some information about physics behind the
1134: horizon. In this approach, the one boundary thermal description is
1135: recovered by tracing over the Hilbert space of the other boundary
1136: CFT as in the real-time or thermofield formalism for thermal field
1137: theory \realtime. The thermal state counting entropy arises from
1138: the entanglement entropy of the pure entangled ``Hartle-Hawking"
1139: state in which expectation values are computed. This realizes an
1140: old idea of Israel \IsraelUR.
1141:
1142: These ideas were explored in some detail for $d=3$, the BTZ black
1143: hole, in \KrausIV. The finiteness of amplitudes (manifest from the
1144: boundary or bulk Euclidean point of view) was shown to arise in
1145: the bulk from $i \epsilon$ regulation of the metric singularity
1146: inherent in analytic continuation from the Euclidean theory and
1147: from cancellations between future and past singularities.
1148:
1149: In addition, \KrausIV\ gave a simple demonstration of how boundary
1150: correlation functions can probe physics behind the horizon by
1151: studying the correlator of two operators, one on each asymptotic
1152: boundary, each creating a large mass bulk particle.
1153: %See Figure
1154: %[Fig of d=3 Penrose diagram with geodesic connecting two boundary
1155: %points].
1156: As the mass $m \rightarrow \infty$, the correlator can be
1157: evaluated in the semiclassical geodesic approximation and is given
1158: by $\exp(-m \len)$. Here $\len$ is the proper length of the
1159: spacelike geodesic joining the boundary points (see Fig.5a). Because
1160: the geodesic passes through spacetime regions inside
1161: the horizon, this boundary correlator
1162: reveals information about the geometry behind the horizon. This
1163: closely parallels a prior calculation done in the ``geon" geometry
1164: \LoukoTP.
1165:
1166: As for all such quantities, an ``outside the horizon"
1167: interpretation
1168: can be given. By tracing the Hartle-Hawking state, perturbed
1169: by the operator on one boundary, over the Hilbert space of the CFT
1170: on that boundary, a perturbed density matrix for the remaining CFT
1171: is created. This density matrix corresponds in the bulk to a
1172: modified boundary condition on the horizon, which can have
1173: nontrivial correlations with an asymptotic boundary field. The
1174: nature of this boundary condition is obscure from the ``outside
1175: the horizon" point of view, but natural when physics behind the
1176: horizon is taken into account.
1177:
1178:
1179: The BTZ geodesics illustrated in Fig.5a pass quite close to the
1180: singularity for moderate boundary times. But the correlation
1181: function is relatively structureless. Presumably this reflects
1182: the simple orbifold geometry of the BTZ black hole. The curvature
1183: of the BTZ geometry is constant except for a delta function at the
1184: singularity.
1185:
1186: Our goal in this paper is to explore further what information from
1187: behind the horizon can be found in boundary AdS/CFT correlators.
1188: To consider a situation with a more interesting geometry, we
1189: focus on AdS Schwarzschild black holes in $d>3$ , and in
1190: particular in $d=5$ where the boundary CFT is four dimensional
1191: ${\cal N }=4$ super Yang Mills theory. The geometry near the
1192: singularity in such black holes approaches that of $d>3$
1193: Schwarzschild, and so the curvature diverges as one approaches the
1194: singularity. This affects the geodesics dramatically.
1195:
1196: We begin in Section 2 by exploring in detail the causal structure
1197: and spacelike geodesics in $d=5$ \sads. (Other dimensions
1198: greater than three behave similarly.) We find several unusual
1199: features. First, the Penrose diagram that encodes the causal
1200: structure of the geometry is not a square, unlike the $d=3$ case.
1201: The lines describing the singularity bend inward toward the center
1202: of the diagram. This allows a nearly null geodesic beginning at
1203: the boundary at a finite nonzero time $t_c$ to ``bounce" off the
1204: singularity and hit the other boundary at the symmetric point.
1205: This behavior, which does not occur for $d=3$, is depicted in
1206: Fig.5b.
1207:
1208: After submitting this paper we have learned that such features have previously
1209: been found in a general study of effectively two dimensional metrics \KloschQV\ \foot{We are grateful
1210: to Jorma Louko for bringing this reference to our attention.} .
1211:
1212: In Section 3 we turn to correlators in the geodesic approximation.
1213: If the bounce geodesic were to dominate, the correlator would
1214: become singular as $t \rightarrow t_c$ because the proper distance
1215: goes to zero as the geodesic becomes null. This would be a kind
1216: of ``light cone singularity." But general considerations about the
1217: boundary field theory rule out this kind of behavior. In fact this
1218: geodesic does {\it not} dominate the correlator. There are in
1219: general multiple geodesics that connect the two boundary points (Fig.12).
1220: At $t=0$ their proper distances coincide, creating a branch point
1221: in the correlator which behaves as $t^{4/3}$ for small $t$. By
1222: studying various resolutions of this branch point, we show that as
1223: $t$ increases from $0$, the correlator defined by the boundary CFT
1224: is given by a symmetric sum of the two complex branches of this
1225: expression. Each of these can be attributed to a complex geodesic
1226: in complexified spacetime. At large $t$ these complex branches
1227: reproduce the correct quasinormal modes for this black hole, which
1228: are complex, unlike the $d=3$ case.
1229:
1230: But the correlator is an analytic function of $t$ and can be
1231: continued onto the real sheet. (This can be done in a
1232: computationally effective manner, as discussed in Appendix F.) On
1233: this sheet, the ``light cone singularity" does appear. So the
1234: boundary correlator {\it does} contain information about the
1235: singularity, albeit in a subtle way.
1236:
1237: In Section 4 we extend the above analysis to include finite $m$,
1238: finite string length $l_s$ ($\alpha '\sim l_s ^2$), and finite
1239: string coupling ($g_s$) effects. We take wrapped D-branes (giant
1240: gravitons \McGreevyCW) as explicit examples of heavy probe
1241: particles. The branch point at $t=0$ is smoothed out for any
1242: finite $m$, preventing analytic continuation to the bouncing
1243: geodesic. But at each order in the $1/m$ expansion the branch
1244: point persists and one can follow this geodesic and the
1245: accompanying fluctuation corrections into the bouncing domain. The
1246: $1/m$ corrections are given by a heat kernel expansion with
1247: coefficients related to the curvature and are large when the
1248: geodesic passes near the singularity. These quantities give a
1249: clear example of information about the singular geometry being
1250: coded into boundary CFT correlators.
1251:
1252: The main corrections at finite but small $l_s$ (with $g_s$ kept
1253: equal to zero) can be expressed as modifications of the
1254: supergravity field equations and hence of the background metric,
1255: dilaton, etc.\ fields. Such small modifications do not change the
1256: basic picture outlined above. The branch point at $t=0$ either
1257: persists or, more generically, splits into two nearby branch
1258: points. In either case we can follow the geodesic onto the real
1259: sheet. Of course, for $t$ near $t_c$, we expect the $l_s$
1260: corrections to become large because the geodesic passes through
1261: regions of large curvature.
1262:
1263: At large $l_s$ the 't Hooft coupling $\lambda$ in the boundary
1264: gauge theory becomes small. In weak coupling perturbative gauge
1265: theory amplitudes are smooth at $t=0$. There is no branch point.
1266: Therefore, for consistency, there must be a phase
1267: transition at finite $\lambda$ (for infinite $m$).
1268:
1269: Finite $g_s$ is a particularly difficult regime in which to study
1270: these phenomena. Wrapped D-brane masses are $\sim 1/g_s$ and so
1271: are finite here. This smooths out the branch point and would seem
1272: to prevent analytic continuation to the $t_c$ singularity. But by
1273: taking appropriate $g_s \rightarrow 0 $ limits we can extract
1274: behavior around the singularity to all orders in $g_s$, as well as
1275: study certain leading nonperturbative effects.
1276:
1277: These results demonstrate that a significant amount of information
1278: from behind the horizon, and in particular from near the
1279: singularity, is encoded in boundary theory correlators. In
1280: Section 5 we discuss the meaning of these results further.
1281:
1282: %------------------------------------------------------------------
1283: \newsec{Classical geometry}
1284:
1285: We will devote this section to the exploration of the classical
1286: Schwarzschild-AdS geometry. As is well known, the \sads\
1287: spacetimes contain spacelike curvature singularities, event
1288: horizons, and timelike boundaries, all of which will play an
1289: important role in the discussion. First, we analyze the causal
1290: structure in detail, arriving at the surprising observation that,
1291: unlike the three dimensional case, for $d>3$ the Penrose diagram
1292: of \sads\ spacetimes cannot be drawn as a square, with both
1293: boundaries and singularities represented by straight lines. Next,
1294: we study spacelike geodesics in this geometry, which, as we will
1295: discuss in the next section, provide a method for approximating
1296: the relevant boundary correlators.
1297:
1298: \sads\ represents a $d \ge 3$ dimensional, two-parameter family of
1299: solutions characterized by the size of the black hole and the AdS
1300: radius. For simplicity we will focus on the large black hole
1301: (planar) limit, where %the horizon radius $r_+$ is much larger than
1302: the AdS radius is much smaller than the horizon radius.
1303: Also, we will from now on concentrate on $d=5$. The geometry
1304: for finite-sized black holes, as well as in other
1305: dimensions $d>3$, is qualitatively similar.
1306: We will contrast this with the quite
1307: different case of the $d=3$ BTZ black hole \KrausIV .
1308:
1309: The metric of the 5-dimensional \sads\ spacetime is given by
1310:
1311: \eqn\metric{\eqalign{ ds^2 &= - f(r) \, dt^2 + {dr^2 \over f(r)} +
1312: r^2 \, d\Om_3^2 \cr f(r) &= {r^2 \over R^2} + 1 - {\rh^2 \over
1313: r^2} \( {\rh^2 \over R^2} + 1 \),}} where $\rh$ is the horizon
1314: radius and $R$ is the AdS radius. We work in the limit of the
1315: infinitely massive black hole $r_+/R \rightarrow \infty$, where
1316: the metric simplifies. In particular, if we rescale the
1317: coordinates $r \rightarrow {\rh \over R} \, r, ~ t \rightarrow {R
1318: \over \rh} \, t$, measure lengths in AdS units so $R=1$, and
1319: suppress the angular coordinates $\Omega$ that will not concern us
1320: here, the metric in the $t$-$r$ plane becomes
1321:
1322: \eqn\metric{\eqalign{ ds^2 &= - f(r) \, dt^2 + {dr^2 \over
1323: f(r)} \, \cr f(r) &= r^2 - {1 \over r^2}\, .}}
1324:
1325: There is a genuine curvature singularity at $r=0$, and the
1326: boundary of the spacetime is approached as $r \to \infty$. The
1327: Schwarzschild coordinates used in \metric\ also have a coordinate
1328: singularity at the horizon $r=1$. This is not a big obstacle,
1329: since one can pass to Kruskal coordinates $(T, X)$ which cover
1330: the full globally extended spacetime. (The Kruskal coordinate
1331: chart is constructed explicitly for this geometry in Appendix A).
1332: However, the global extension can be discussed more conveniently
1333: for our purposes by using four Schwarzschild coordinate patches
1334: (corresponding to the two asymptotic regions (I and III) with
1335: $r>1$, plus the regions inside the black hole (II) and the white
1336: hole (IV) with $0<r<1$; cf.\ Fig.1).
1337:
1338: \ifig\figSchcoords{Complexified coordinates for the AdS black
1339: hole: the time coordinate is complex in the extended spacetime,
1340: but with constant imaginary part in each wedge, as indicated
1341: above. The wedges are separated by the horizons at $r=1$, where
1342: $t$ diverges. Note that the real part of $t$ increases upward in
1343: wedge I, downward in wedge III, and to the right (left) in wedge
1344: II (IV).} {\epsfxsize=7cm \epsfysize=5.4cm \epsfbox{fig_newsch.eps}}
1345:
1346: These four patches can be embedded in complexified
1347: Schwarzschild time, \eqn\compltime{t= t_L +i \, t_E}
1348: where $t_L$ and $t_E$ denote times on
1349: the Lorentzian and Euclidean slices respectively. In each
1350: coordinate patch, $t$ has a constant imaginary part, as indicated
1351: in \figSchcoords; we will define this to be 0 for the right
1352: asymptotic region. Crossing a horizon shifts the imaginary part by
1353: $i \, \beta/4$. Setting $t=i \, t_E $ produces Euclidean \sads\
1354: with a periodic imaginary time coordinate $t_E$. The period of
1355: $t_E$ is $\beta$, the inverse temperature $T$ of the black hole.
1356: One can move a point from the boundary of region I to the
1357: symmetric point on the boundary of region III in \figSchcoords\
1358: by rotating by half a period; that is, shifting $t \rightarrow -t
1359: - i \, \beta /2$. The minus sign accounts for the opposite
1360: direction of time in region III. In the geometry \metric ,
1361: $\beta=\pi$. The CFT correlators can be defined for complex time
1362: and are analytic, so such an extension is natural from the
1363: boundary side. The periodicity is just that of finite temperature
1364: field theory.
1365:
1366: % The remainder of this section is structured as follows:
1367: % We will first study radial null geodesics, because these reflect
1368: % the causal structure captured by the Penrose diagram. In
1369: % particular, we know that in a spherically symmetric spacetime,
1370: % radial null rays will always be 45-degree lines in the Penrose
1371: % diagram.
1372: % % We will first motivate the unusual shape of the Penrose diagram by
1373: % % considering radial null geodesics in the spacetime.
1374: % Having addressed the geometry, we will proceed to discuss the properties
1375: % of spacelike geodesics on this spacetime.
1376:
1377:
1378:
1379: %------------------------------------------------------------------
1380: \subsec{Radial null geodesics}
1381:
1382:
1383: We now turn to a major focus of this paper, the study of geodesics
1384: in the geometry \metric . We consider a radial ($\Omega$
1385: constant) path, $x = (t(s), r(s))$, where $s$ is an affine
1386: parameter. Geodesic paths can be found by extremizing the action
1387: \eqn\geoact{S = \int ds ~{\dot x}^2,} where ${\dot x = {d x \over
1388: d s}}$. We can then find a unit speed parametrization by requiring
1389: ${\dot x}^2 =0, +1, -1$ for null, spacelike, and timelike
1390: geodesics, respectively.
1391:
1392: Due to the symmetry described by the Killing field $\dt$, there
1393: exists a conserved quantity $E$ along each geodesic, \eqn\Ecm{ E =
1394: \td \, f(r). } To map out causal structure we first consider null
1395: geodesics. In a spherically symmetric spacetime, radial null rays
1396: will always be 45-degree lines in the Penrose diagram. Here ${\dot
1397: x} ^ 2 = -f (r) \, {\td}^2 + {{\rd}^2 \over f(r)} = 0$, which when
1398: combined with \Ecm\ becomes \eqn\rdotsq{ \rd^2 = E^2~ .}
1399:
1400: Let us consider the coordinate time $t(r)$ along an ingoing radial
1401: null geodesic which starts from the boundary $r = \infty$ at
1402: $t=t_0$. From \Ecm\ and \rdotsq\ we have that
1403:
1404: \eqn\tra{ t(r) = \int {\td \over \rd'} \, dr' = t_0 +
1405: \int^{\infty}_{r} {dr' \over f(r')}.}
1406: For geodesics which cross the
1407: horizon, integrating over the pole at $r = 1$ gives a constant
1408: imaginary part $-i {\pi \over 4}$, so that for $r < 1$, \eqn\tre{
1409: t(r) = t_0 - {1 \over 2} \[ \tan^{-1} r - {\pi \over 2} -
1410: \tanh^{-1} r
1411: \] - i \, {\pi \over 4}. } The the last term is simply $i \beta
1412: /4$, where again $\beta = 1/T = \pi$ is the Bekenstein-Hawking
1413: temperature (cf.\ \figSchcoords) for this geometry.
1414:
1415: This result implies that a null geodesic which starts at the
1416: boundary at $t_0 = 0$ reaches the singularity $r=0$ at
1417: $ %\eqn\tsing{
1418: t(0) = {\pi \over 4} \( 1 - i \)$. %}
1419: Since the real part of $t(0)$ is positive,
1420: %from \figSchcoords\ this means that
1421: the geodesic hits the singularity {\it off-center}
1422: in the Penrose diagram; in fact it hits closer to the right
1423: boundary, where it originated.
1424: %\foot{
1425: %Technically, %this would have to be defined a bit more rigorously,
1426: %since the singularity is not a part of our spacetime, we really mean that
1427: %they come arbitrarily close (metrically) near enough to the singularity.};
1428: %the previous results imply that they do not.
1429:
1430: \ifig\figPdg{Behaviour of radial null geodesics suggests the shape
1431: of the Penrose diagram (wavy lines represent the singularities,
1432: dashed lines the horizons, vertical lines/curves the boundaries,
1433: and the solid 45 degree lines the radial null geodesics): a) if
1434: geodesics meet at the singularity, the diagram can be drawn as a
1435: square; if they don't meet, b) with straight boundaries, the
1436: singularities are bowed in, or c) with straight singularities, the
1437: boundaries must be bowed out.} {\epsfxsize=12cm \epsfysize=3.7cm
1438: \epsfbox{fig_Pdg.eps}}
1439:
1440: This immediately suggests the nature of the Penrose diagram, as
1441: illustrated in \figPdg. If the Penrose diagram were a square, as
1442: in \figPdg a, then radial null geodesics sent off from the
1443: boundary at $t=0$ (defined as the horizontal line of symmetry of
1444: the Penrose diagram) would have to meet at the same point on the
1445: singularity. On the other hand, if the geodesics don't meet at or
1446: before reaching the singularity, then the
1447: singularity on the Penrose diagram must be drawn {\it bowed
1448: in}, assuming we draw the boundaries straight, as in \figPdg b.
1449: Alternately, if the singularity were to be
1450: represented by a straight line, the boundaries would have to be
1451: bowed out, as in \figPdg c, in order to reproduce the calculated
1452: behaviour of the geodesics. Note that the Penrose diagrams
1453: sketched in \figPdg b and \figPdg c are equivalent, since we can
1454: use a conformal transformation to straighten out the singularities
1455: of \figPdg b, but at the expense of bowing out the boundaries.
1456: This will be discussed in greater detail in Appendix B, where the
1457: Penrose diagram for this spacetime is constructed explicitly.
1458: Such Penrose diagrams have previously been discussed in \KloschQV\ .
1459:
1460: Conversely, we can ask at what time does a null geodesic have to
1461: start from the boundary in order to reach the singularity at
1462: $\Re[t(r=0)] = 0$, i.e.\ in the ``middle'' of the singularity.
1463: This will play an important role later. Denoting such a time by
1464: $\tc$, we see that \eqn\tcrit{ \tc = - {\pi \over 4} = - {\beta
1465: \over 4} .} In general dimension $d$, one finds that $t_c = - \pi
1466: \( (d-1) \, \tan {\pi \over (d-1)} \)^{-1}$, which vanishes only for
1467: $d=3$. Thus we see that, with the exception of the $d=3$ BTZ black
1468: hole, all singularities in the Penrose diagrams of \sads\ are
1469: bowed in as in \figPdg b.
1470:
1471: \ifig\krusk{Radial null geodesics on Kruskal diagram, a) for
1472: $d=3$, and b) for $d>3$ (radial null lines lie at 45 degrees, the
1473: horizons are represented by dashed lines, and the boundaries and
1474: singularities by the solid hyperbolas). In order for radial null
1475: geodesics to meet at the singularity at $X=0$, they have to start
1476: at a) $t=0$ for the $d=3$ case, and b) $t = - t_c < 0$ for the
1477: $d>3$ case. } {\epsfxsize=12cm \epsfysize=6.3cm
1478: \epsfbox{krusk1.eps}}
1479:
1480: In Kruskal coordinates $(T, X)$, the future and past singularities
1481: of the black hole are the hyperbolas $T^2 - X^2 = 1$, while the
1482: AdS boundaries are given by $T^2 - X^2 = - e^\pi$. (By contrast,
1483: in the $d=3$ BTZ geometry, the singularities are at $T^2 - X^2 =
1484: 1$ and the boundaries are at $T^2 - X^2 = - 1$.) The curved nature of
1485: the boundary of the Penrose diagram is reflected in these
1486: coordinates by the asymmetry in the radii of the two hyperbolas.
1487: This asymmetry allows a null geodesic starting at $t_c$ to
1488: ``bounce off" the singularity. This is illustrated in \krusk ,
1489: where the $d=3$ case (\krusk a) is contrasted with the
1490: higher-dimensional case (\krusk b).
1491: In Appendices A and B we explain this in further detail, and construct
1492: the Penrose diagram explicitly. (The latter will be used in the
1493: next subsection to produce the Penrose diagrams in Fig.5.)
1494: These issues are also discussed in \KloschQV\ .
1495:
1496: %------------------------------------------------------------------
1497: \subsec{Radial spacelike geodesics}
1498:
1499: % \ifig\figgeods{Plot of the past-directed radial symmetric geodesics.
1500: % The spacelike geodesics converge to a ``high-energy" null geodesic
1501: % which comes arbitrarily close to the singularity.
1502: % By symmetry, there is of course a ``reflected" family of future directed
1503: % geodesics of the same type.}
1504: % {\epsfxsize=7cm \epsfysize=7cm \epsfbox{penrose5.eps}}
1505:
1506: % \ifig\adsthree{Symmetric spacelike geodesics of the BTZ black
1507: % hole. Due to the fact that only finite $E$ is required to come
1508: % arbitrarily close to the singularity, none of these geodesics is
1509: % nearly null.} {\epsfxsize=7cm \epsfysize=7cm
1510: % \epsfbox{ads3geo.eps}}
1511:
1512:
1513: Let us now turn from the qualitative structure of the Penrose
1514: diagram to quantities which we can probe from the CFT side: in
1515: particular, the proper length along spacelike geodesics. In this
1516: subsection, we will consider purely radial spacelike
1517: geodesics.\foot{For completeness, we
1518: present in Appendix C an analogous treatment
1519: for spacelike geodesics carrying angular momentum. For the
1520: remainder of the paper, however, we need only the results about
1521: geodesics with zero angular momentum.}
1522: The spacelike geodesics satisfy ${\dot x} ^ 2 = -f (r) {\td}^2 + {{\rd}^2 \over
1523: f(r)} = 1$, giving \eqn\spgeod{ E = \td \, f(r), \ \ \ \ \ \ \rd^2 -f(r)
1524: = E^2 .} Here $E$, as before, is a conserved quantity resulting
1525: from the time translation symmetry of the geometry. We can think
1526: of it as the total energy of a particle moving in a potential
1527: $V(r) = -f(r)$ (where the minus sign is due to the fact that we
1528: are considering spacelike geodesics).
1529:
1530: \ifig\energy{The ``potential energy" in the mechanics problem for
1531: spacelike geodesics in \sads\ in $d=3$ and $d=5$. The potential
1532: diverges at $r=0$ for all $d > 3$.} {\epsfxsize=7cm
1533: \epsfysize=5.9cm \epsfbox{energy.eps}}
1534:
1535: In \energy, we plot the potential $V(r)$ for $d=3$ and $d=5$
1536: \sads. In the BTZ case, the energy required to reach $r=0$ is
1537: finite. As a result, the proper time along the geodesic is large.
1538: By contrast, in the higher dimensional examples the potential
1539: diverges at the singularity, so that large $E$ is required to
1540: approach $r=0$. At large $E$ the particle is moving fast and so
1541: very little proper time is covered. The result as $E \rightarrow
1542: \infty $ is a geodesic that is null everywhere except in a
1543: vanishingly small region very near the singularity where it
1544: ``bounces" off. As explained above, this occurs at $t_0 = t_c$.
1545:
1546: \ifig\figgeods{Symmetric spacelike radial geodesics of the \sads\
1547: black hole, plotted on the corresponding Penrose diagrams. a)
1548: $d=3$: Because only finite $E$ is required to come arbitrarily
1549: close to the singularity, none of these geodesics is nearly null.
1550: b) $d>3$: the spacelike geodesics converge to a large $E$ almost
1551: null geodesic which comes arbitrarily close to the singularity. }
1552: {\epsfxsize=14cm \epsfysize=7.5cm \epsfbox{fig_Pdgeods.eps}}
1553:
1554: The behaviour of radial spacelike symmetric geodesics (i.e. those
1555: which reach their minimum value of $r$ at $t= {i \beta \over 4}$)
1556: in $d=3$ and $d=5$ is plotted on the corresponding Penrose
1557: diagrams in \figgeods. As explained above, in the $d=3$ BTZ case
1558: (\figgeods a), the geodesics do not show any striking feature as
1559: the starting time $t_0$ is varied. On the other hand, in higher
1560: dimensions such as $d=5$ (\figgeods b), spacelike symmetric
1561: geodesics exist only for a finite range of starting times, $t_0
1562: \in (-t_c, t_c)$. For starting times outside this range, no such
1563: geodesics exist. This is in accord with the expectation that the
1564: genuine curvature singularity of the higher dimensional black
1565: holes has a pronounced effect on geodesics in its vicinity, as
1566: opposed to the $d=3$ BTZ orbifold singularity.
1567:
1568: Let us first consider the $E=0$ case. From \spgeod, we see that
1569: $\td = 0$; $t(s) = 0$ is the only consistent solution. We have
1570: that $\rd^2 = f(r)$. Such a geodesic crosses the Penrose diagram
1571: in the middle, so that it goes from $r=\infty$ at one boundary,
1572: down to the horizon $r = 1$, and back out to $r=\infty$ at the
1573: other boundary. Since it never penetrates inside the horizon,
1574: this geodesic can equally well be embedded in Euclidean \sads.
1575: It is the geodesic that crosses over the tip of the cigar to the
1576: antipodal point $t_E = \beta/2$ on the Euclidean time circle (Fig.
1577: 8 ).
1578:
1579: The proper length along such a geodesic is then given by
1580: \eqn\length{ \len = 2 \int_{1}^{\rinf} {dr \over \rd} = 2
1581: \int_{1}^{\rinf} {dr \over \sqrt{f(r)} }} where $\rinf \to \infty$
1582: is the large $r$ regulator standard in AdS/CFT \AharonyTI\ .
1583: Integrating this explicitly, we have \eqn\Lresc{
1584: \len % = 2 \int_{1}^{\rinf} \sqrt{{r^2 \over r^4 -1 }} \, dr
1585: = \ln \( \rinf^{2} + \sqrt{ \rinf^4 -1} \) \approx 2 \ln \rinf +
1586: \ln 2 + \CO(\rinf^{-4}) } so that taking $\rinf \to \infty$ and
1587: subtracting off the universal logarithmically divergent piece, we
1588: obtain the regularized proper length \eqn\Lreg{ \len_{\rm reg} = \ln
1589: 2 } In the future we will drop the subscript on $\lenr$.
1590: %In passing, we note that
1591: %we can easily generalize this to $d$ dimensions, where
1592: %\eqn\Lrescd{
1593: %\len = 2 \int_{1}^{\rinf} \sqrt{{r^{d-3} \over r^{d-1} -1 }} \, dr
1594: %\approx 2 \ln \rinf + {4 \over d-1} \, \ln 2
1595: % + {1 \over (d-1)} \rinf^{-(d-1)} + ... }
1596: %which yields the regulated length
1597: %\len_{\rm reg} = {4 \over d-1} \, \ln 2 }
1598: %dimension.
1599:
1600: Let us now consider what happens as we increase the energy $E$.
1601: From \spgeod\ we see that the geodesic now penetrates a finite
1602: distance inside the horizon. Specifically, let $\ri$ denote the
1603: smallest value of the radial coordinate reached along such a geodesic. This
1604: corresponds to the
1605: classical turning point of the particle motion, which is
1606: obtained by solving \eqn\rieqn{ E^2 + f(\ri) = 0 ~.} Since $f(r)$
1607: is positive outside the horizon and negative inside, $\ri \le 1$.
1608: For very high energy $E \gg 1$, the geodesic comes very close to
1609: the singularity, $\ri \ll 1$, so that $f(r) \sim - {1 \over r^2}$,
1610: which means that $\ri \sim {1 \over E}$.
1611:
1612: In general, spacelike geodesics with arbitrary $E$ and starting
1613: time $t_0$ will not be symmetric. We can specialize to this case,
1614: however, and obtain all others by $t$ translation. For a given
1615: $E$, there exists a starting time $t_0$ on the boundary in region
1616: I for which the geodesic is symmetric. In particular, for a
1617: symmetric geodesic the turning point $r = \ri$ has to occur at
1618: $\Re[t(\ri)] = 0$. The turning point must be in region II so
1619: $\Im[t(\ri)]= -\beta/4$. The geodesic ends in region III with
1620: coordinate $t= -t_0 - i \beta/2$. As mentioned before, the real
1621: part has a minus sign because the Schwarzschild coordinate time
1622: runs in the opposite directions in regions I and III. The geodesic
1623: equations give \eqn\tstartsp{ t_0-t(\ri)=t_0+ i \beta/4 = -
1624: \int_{\ri}^{\infty} {\td \over \rd} \, dr
1625: = - \int_{\ri}^{\infty} {E \over f(r) \, \sqrt{ E^2 +
1626: f(r) }} \, dr. } For $E=0$, we recover $t_0=0$. One can also check
1627: that as $E \to \infty$, $t_0 \to \tc$; and in fact, for large
1628: energies one can see from examining the integral that \eqn\tdiff{
1629: (t_0 - \tc) \sim { 1 \over E }.} For each such geodesic there
1630: exists a mirror one obtained by taking $t \rightarrow -t, ~ E
1631: \rightarrow -E$.
1632:
1633: The proper length, analogous to \length, along a symmetric
1634: spacelike geodesic originating at $\rinf$, going down to $\ri$,
1635: and then back out to $\rinf$ in the other asymptotic region, is
1636: \eqn\lengthE{ \len = 2 \int_{\ri}^{\rinf} {dr \over \sqrt{E^2 +
1637: f(r)} }\ .} Upon evaluating and regularizing this integral, we
1638: obtain \eqn\lengthEr{\len = \ln \({2 \over \sqrt{{E^4} / 4 + 1}}\)
1639: .} As $E$ goes to infinity the geodesic looks more and more null,
1640: and so at any fixed radial cutoff its proper length goes to zero.
1641: This is apparent already from \figgeods b, where we can see
1642: explicitly that symmetric
1643: spacelike geodesics in the $E \to \infty$ limit approximate the
1644: null geodesic which bounces off the singularity.
1645: The regularized length along an $E \to \infty$ geodesic
1646: (which is roughly the difference between
1647: its proper length and that of the $E = 0$ geodesic) goes to
1648: negative infinity at $t_c$,
1649:
1650: \eqn\lengthErapp{\len \sim 2 \ln \(t-t_c\).}
1651: We will make use of this result in the next section, after we discuss
1652: the correlation functions.
1653:
1654: % only the past-directed spacelike symmetric geodesics are shown.
1655: % For each such geodesic, there is a future-directed mirror geodesic
1656: % which approaches the future singularity.
1657: % \foot{For completeness, we
1658: % present in appendix C an analogous treatment
1659: % for spacelike geodesics carrying angular momentum. For the
1660: % remainder of the paper, however, we need only the results about
1661: % geodesics with zero angular momentum}
1662:
1663:
1664:
1665: %------------------------------------------------------------------
1666: \newsec{Correlation functions}
1667:
1668: According to the AdS/CFT correspondence, \sads\ in $d=5$ is dual
1669: to
1670: ${\cal N}=4$ SYM theory at finite temperature. % We will
1671: % concentrate on the case of infinite mass \sads\ (corresponding to
1672: % infinite temperature, where the CFT lives in infinite volume).
1673: As before, we will consider the large black hole limit, so the CFT
1674: effectively lives in infinite volume. Standard CFT finite
1675: temperature boundary correlators are dual to bulk correlation
1676: functions with insertions all in one asymptotic region (e.g.,
1677: region I). There are other boundary correlators one might want
1678: to study. For example, one could put an operator on the boundary
1679: of region I, and another on the boundary of region III
1680: \refs{\MaldacenaKR , \KrausIV} . We have discussed the geodesics
1681: connecting such points. In the field theory, such correlators
1682: have a natural representation in the thermofield double or
1683: real-time description of finite temperature field theory
1684: \realtime.
1685:
1686: In the real-time formalism, one takes the tensor product of two
1687: copies of the original field theory labeled by $1,2$. The two
1688: copies are decoupled, and the total Hamiltonian is \eqn\Hdiff{
1689: H_{tf} \equiv H \otimes I - I \otimes H^*, } where $H$ is the
1690: Hamiltonian for the original theory. We now construct the
1691: entangled state \eqn\state{ |\psi \rangle= {1 \over Z^{\12}}
1692: \sum_i e^{-\12\beta E_i}|E_i,E_i\rangle, } where
1693: $|E_i,E_j\rangle=|E_i\rangle\otimes|E_j\rangle$, and $|E_i\rangle$
1694: are energy eigenstates. The state $|\psi\rangle$ is a particular
1695: eigenvector of $H_{tf}$ with eigenvalue zero. Correlations
1696: between subsystems $1$ and $2$ are due to the entanglement in
1697: $|\psi \rangle$.
1698:
1699: Operators which belong to subsystem $1$ have the form $A \otimes
1700: I$ (where $I$ is the identity operator), and will be denoted
1701: $A_1$. Operators associated with subsystem $2$ are defined in a
1702: similar manner, except with an additional rule of hermitian
1703: conjugation: \eqn\atwo{ A_2 \equiv I \otimes A^{\dag}. } Standard
1704: thermal correlation functions may be written as an expectation
1705: value: \eqn\tfcora{ \langle \psi | A_1(0) B_1(t) | \psi \rangle .}
1706: As can be easily seen from the form of $ | \psi \rangle$, \tfcora\
1707: is simply the thermal expectation value of $A(0) B(t)$, evaluated
1708: in a thermal density matrix at inverse temperature $\beta$. The
1709: state counting entropy observed in subsystem $1$ is the entropy of
1710: entanglement of the state $|\psi \rangle$. In field theory, no
1711: physical significance is usually attached to correlators involving
1712: both subsystems, but we can certainly define them; for example
1713:
1714: \eqn\tfcorb{ \langle \psi| A_1(0) B_2(t) |\psi\rangle. }
1715: In the finite temperature AdS/CFT correspondence, \tfcorb\ has a
1716: simple interpretation \MaldacenaKR : it corresponds in the bulk to
1717: a correlator between operators on the two disconnected boundaries
1718: of the spacetime.
1719:
1720: It is not hard to see that one can compute \tfcorb\ by
1721: analytically continuing \tfcora.
1722:
1723: \eqn\tfcorc{ \langle \psi| A_1(0) B_2(t) |\psi\rangle = \langle
1724: \psi | A_1(0) B_1(-t-i\beta/2) | \psi \rangle . }
1725: This is the analog of using complexified Schwarzschild time as
1726: discussed in Section 2. We see that these two sided correlators
1727: are just part of the information contained in ordinary thermal
1728: correlators, as a function of complex time.
1729:
1730:
1731: We can now use our knowledge of geodesics in \sads\ to study
1732: correlation functions in the CFT. The prescription is to
1733: determine the 2-point CFT correlators via AdS/CFT from a
1734: computation of the bulk propagator. We assume at this point that
1735: we have a scalar field of mass $m$ in the bulk, dual to some
1736: operator in the CFT whose 2-point function we want to compute. The
1737: bulk propagator is given by a sum over paths between two points in
1738: the bulk, with each path contributing $e^{-m \len}$, $\len$ being
1739: the proper length of the path, suitably regulated. In the limit
1740: of large $m$, this sum will generically be dominated by the
1741: shortest geodesic connecting the two points. So the 2-point CFT
1742: correlator should go like $e^{-m \len}$ (up to $1 \over m$
1743: corrections), where $\len$ is the regularized length of the
1744: shortest geodesic between the two points\foot{This approach and
1745: some of the pitfalls involved are discussed in \LoukoTP.}.
1746:
1747: In section 3.1 we show that a naive application of this
1748: prescription predicts a ``light cone'' pole in the 2-point
1749: opposite side correlator as $ t \to t_c$. However, simple
1750: arguments in the field theory show that this prediction cannot be
1751: correct, at least in the theory defined by analytic continuation
1752: from Euclidean space. In Section 3.2 we address this puzzle by
1753: examining the branch cut singularity in $\len$ at $t=0$ that is
1754: due to a coalescence of geodesics. We first resolve the branch
1755: cut by taking $\rinf $ large but finite. We find that the
1756: unambiguously correct geodesic in Euclidean space bifurcates into
1757: two complex geodesics at small Lorentzian time $t_L$. These
1758: geodesics must determine the correct answer. In section 3.3 we
1759: elucidate this further by moving this bifurcation onto the
1760: Euclidean section. This involves taking the black hole mass large
1761: but finite. Euclidean correlators are unambiguously determined by
1762: the shortest geodesics at large $m$, and so we can determine the
1763: correct choice of geodesics through the bifurcation. The dominant
1764: geodesics continue to the complex ones mentioned above as one goes
1765: to the Lorentzian section. However, at large $m$ it is possible to
1766: analytically continue correlators around these bifurcations by
1767: following geodesics that are no longer dominant. This allows us to
1768: find the $t_c$ singularity on a secondary sheet of the branched
1769: correlation function. This $t_c$ singularity gives information
1770: about the black hole singularity. It shows that such information
1771: is encoded in CFT correlators, albeit in a subtle way involving
1772: analytic continuation. In section 3.4 we introduce a simple model
1773: of the bifurcation given by an ordinary integral. The properties
1774: of this model can be determined precisely and agree with the
1775: picture we have obtained. We demonstrate here that all orders
1776: fluctuation corrections around a given saddle point can be
1777: analytically continued as well. This indicates that additional
1778: information about the black hole singularity is also accessible.
1779: In section 3.6 we show that our results for the correlator agree
1780: with the known values for the quasinormal mode frequencies of the
1781: \sads\ black hole. This gives independent confirmation of our
1782: analysis.
1783:
1784: %------------------------------------------------------------------
1785: \subsec{A light cone singularity }
1786:
1787: Recall from Equation \lengthErapp\ that the regularized length of
1788: a symmetric geodesic with boundary time $t$ diverges
1789: logarithmically at $t_c$. Therefore the correlation function
1790: behaves like \eqn\lcsing{e^{-m \len} \sim {1 \over (t -
1791: t_c)^{2m}}\ .} We seem to have a pole of order $2m$ in the
1792: two-sided
1793: CFT correlator coming from the ``almost null'' % ``lightlike"
1794: geodesic connecting the boundary points.
1795:
1796: This immediately leads to a puzzle. We know that the real-time
1797: two-point correlation function, evaluated in a thermal state with
1798: inverse temperature $\beta$, is \eqn\twopoint{\eqalign{ \langle
1799: \phi(t) \phi(-t) \rangle_\beta = \sum_n e^{- \beta E_n} \langle
1800: E_n | e^{- i H t} \phi(0) e^{ i H t} e^{ i H t} \phi(0) e^{- i H
1801: t}| E_n \rangle \cr = \sum_{n,m} e^{ -\beta E_n - 2 i t (E_n -
1802: E_m)}
1803: \left| \phi_{nm} \right|^2 ,}}
1804: where $\phi_{nm} \equiv \langle E_n | \phi(0) | E_m \rangle$ and
1805: we have assumed $\phi^{\dag} = \phi.$ To get the two sided
1806: result, we can either use the real-time formalism and the state $|
1807: \psi \rangle$, or continue one of the times from the real value
1808: $t$ to $t - i \beta /2$: \eqn\twopointagain{ \langle \psi |
1809: \phi_2(t) \phi_1(-t) | \psi \rangle = \langle \phi(t + i \beta/2 )
1810: \phi(-t) \rangle_\beta = \sum_{n,m} e^{ - {\beta \over 2} (E_n +
1811: E_m) - 2 i t(E_n - E_m)} \left| \phi_{nm} \right|^2. } Note that
1812: the terms in the sum are real and positive for $t=0$; therefore
1813: \eqn\ineq{\left| \langle \phi_2(t) \phi_1(-t) \rangle_\beta
1814: \right| \leq \langle \phi_2(0) \phi_1(0) \rangle_\beta} for any
1815: Hermitian operator $\phi$. However, $\langle \phi_2(0) \phi_1(0)
1816: \rangle_\beta$ can be computed unambiguously in Euclidean space,
1817: using the geodesic approximation. It is certainly finite (since
1818: the two points are on opposite sides of the Euclidean cigar). Thus
1819: we cannot have a singularity in the opposite side correlator at
1820: any time, including $t = t_c$.
1821:
1822:
1823: %-------------------------------------------------------------------
1824: \subsec{Tracking geodesics}
1825:
1826:
1827: We begin exploring this puzzle by computing the boundary time as a
1828: function of $E$ from the integral \tstartsp . The computation is
1829: detailed in Appendix D; the result is
1830:
1831: \eqn\longeqn{\eqalign{t &=
1832: {\half} \ln {\({{{\half} E^2 - E + 1} \over {\sqrt{1 + {\quarter}
1833: E^4}}}\)} - {\half} i \ln {\({{- {\half} E^2 + i E + 1} \over
1834: {\sqrt{1 + {\quarter} E^4}}}\) } \cr \len &= \ln \({2 \over
1835: \sqrt{1 + {\quarter} E^4}}\) ~ ,}} where formula \lengthEr\ has been
1836: repeated for convenience. We have not restricted ourselves to
1837: symmetric geodesics here. The time $t$ is the difference between
1838: $t_f$ and $t_i$, the final and initial boundary Schwarzschild
1839: times; $t=t_f-t_i + i \beta/2$. We have chosen the origin so that
1840: $t=0$ describes endpoints in regions I and III at zero time. For
1841: symmetric geodesics $t=-2 \, t_0$.
1842:
1843: These equations determine $\len (t)$ and hence the large $m$
1844: correlator $\langle \phi \phi \rangle (t) = \exp(-m \len(t))$.
1845: They are central to our analysis.
1846:
1847: Expanding \longeqn\ around $E=0$ we find \eqn\Eexpans{\eqalign{ t
1848: &\sim E^3 \cr \len &\sim - E^4 \cr \len &\sim -t^{\ft}~.}} The
1849: branch point at $t=0$ signals novel behavior. There are three
1850: branches for $\len$ and hence for the correlator at each $t$. It
1851: will turn out the gauge theory chooses a symmetric combination of
1852: the two complex branches. The $t_c$ singularity lies on the real
1853: branch. But knowing the correlator on the complex branch will
1854: allow us to study the real branch by analytic continuation.
1855: Information about the black hole singularity is encoded in the
1856: gauge theory in this manner. The analytic structure of the
1857: Riemann surface defined by \longeqn\ is discussed in Appendix E.
1858:
1859: \ifig\cubic{Plot of $t$ as a functions of $E$, normalized by
1860: $t(E \to \infty)$, for various values of $\rinf$:
1861: the solid line has $\rinf = 2$, while
1862: the dashed line represents $\rinf = \infty$}
1863: {\epsfxsize=7cm \epsfysize=3.2cm \epsfbox{cubic.eps}}
1864:
1865: \ifig\threegeods{Three ``zero time'' geodesics between the two
1866: cutoff boundary points ($t=0, r=\rinf$). The curved dashed lines are
1867: $r = \rinf$.}
1868: {\epsfxsize=6cm \epsfysize=6cm \epsfbox{3geo.eps}}
1869:
1870: To track these branches, it is useful to resolve the branch point.
1871: To do this we take $\rinf$ finite. In the gauge theory this
1872: corresponds to a finite UV cutoff. Equations \Eexpans\ become
1873: \eqn\Eexpanstwo{\eqalign{ t & \sim E^3 - a E \cr \len & \sim - E^4
1874: - b E^2 ~.}} Here $a, b > 0$ are constants that go to zero as
1875: $\rinf \rightarrow \infty$. At time $t=0$ there are now three real
1876: solutions for $E$
1877: and $\len$ (see \cubic). %[DRAWING OF $E^3 + a E$]
1878: These correspond to the three geodesics shown in \threegeods.
1879:
1880:
1881: We can most precisely define gauge theory correlators by
1882: evaluating them in Euclidean time and then analytically continuing
1883: to general time. To compute the $t=0$ two sided correlator this
1884: way we smoothly shift one point of a coincident point Euclidean
1885: correlator by a half period, $i \beta/2$. At $t=0$ we can do the
1886: geodesic computation entirely in the Euclidean \sads\ geometry.
1887: The correct geodesic is clearly the $E=0$ solution of \Eexpanstwo\
1888: which is described in \length .
1889:
1890: Now let us go to small Lorentzian time by increasing the real part
1891: of $t$ by a small amount. The correct geodesic here is certainly
1892: the small deformation of the $E=0$ solution with $E \sim t/a$.
1893: This corresponds to following the central branch of the cubic
1894: curve in \cubic. But at a time $t \sim a^{3/2}$ this root of the
1895: cubic annihilates with another root at the local maximum shown in
1896: \cubic. These solutions then become complex. Complex values of
1897: $E$ correspond to geodesics in the complexified \sads\ spacetime,
1898: where both $r$ and $t$ are complex. These can be described as the
1899: solutions of the equations of motion of the mechanics problem
1900: with complex energy $E$.
1901:
1902: Therefore the gauge theory answer for the correlator, defined by
1903: continuation from Euclidean space, must become complex after a
1904: certain time. The complex branches of \longeqn\ do not contain
1905: the $t_c$ singularity, and so this analysis explains its absence
1906: in the gauge theory. On the complex branches the second derivative of the real part of
1907: $\len(t)$ is negative at $t=0$. In other words, the correlation
1908: function computed by following this branch will decrease as we
1909: move away from $t=0$, consistent with the general field theory
1910: result.
1911:
1912:
1913: %-------------------------------------------------------------------
1914: \subsec{Finite mass black hole}
1915:
1916: To understand the pattern of geodesics better, we now describe how to move the
1917: bifurcation point to Euclidean time. The advantage of studying
1918: the Euclidean correlator is that, in the sum over paths
1919: representation, the weight factor is always positive and so we can
1920: reliably predict that the geodesics with the smallest $\len$
1921: dominate. To move the bifurcation we consider the finite mass
1922: \sads\ geometry and compute $t(E)$ for its geodesics. The result
1923: for small $E$ takes the form (Appendix D):
1924:
1925: \eqn\Eexpansthree{\eqalign{ t &\sim E^3 - a \, E \cr \len &\sim -
1926: E^4 - b \, E^2 ~.}} Here $a < 0$, unlike the finite $\rinf$
1927: resolution described by \Eexpanstwo. The parameters $a$ and $b$ go
1928: to zero as the black hole mass becomes infinite. The parameter
1929: $b$ is still positive.
1930:
1931: If $t= t_L + i \, t_E$ is Euclidean ($t_L = 0$),
1932: then $E$ is pure imaginary. Let $E = - i \, {\tilde E}$,
1933: with ${\tilde E}$ real. Equation \Eexpansthree\ becomes
1934:
1935: \eqn\Eexpansfour{\eqalign{ t_E &\sim {\tilde E}^3 - |a| \, {\tilde E}
1936: \cr \len &\sim - {\tilde E}^4 - |b| \, {\tilde E}^2 ~.}}
1937:
1938:
1939: \ifig\euclidgeo{Geodesics of finite mass Euclidean \sads\ with
1940: finite $\rinf$. The ${\tilde E}=0$ geodesic goes over the top and
1941: has larger $\len$ than the two geodesics with nonzero ${\tilde
1942: E}$.} {\epsfxsize=7cm \epsfysize=7cm \epsfbox{euclidgeo.eps}}
1943:
1944: This cubic is also described by \cubic\ with the horizontal axis
1945: now being ${\tilde E}$. At large ${\tilde E}$, the extension of
1946: \longeqn\ gives $t_E \sim \beta/2$ describing almost coincident
1947: Euclidean points. As ${\tilde E}$ is decreased, entering the
1948: domain of validity of \Eexpansfour, a single solution exists,
1949: describing a single geodesic. But at a particular value of $t_E
1950: \sim |a|^{3/2}$ corresponding to the local maximum of the cubic in
1951: \cubic, a pair of geodesic solutions is created. From
1952: \Eexpansfour\ and the sign of $b$ it is clear that these new
1953: geodesics have larger $\len$ than the original one, so they are
1954: exponentially subdominant. This situation persists until $t_E
1955: =0$, when the geodesic endpoints are antipodal. Here the three
1956: solutions of \Eexpansfour\ are ${\tilde E} = 0, {\tilde E} = \pm
1957: \sqrt{|a|}$. The two geodesics with ${\tilde E} = \pm \sqrt{|a|}$
1958: have the same $\len$, which is less than that of the ${\tilde E} =
1959: 0$ geodesic. This situation is illustrated schematically in
1960: \euclidgeo\ where we have taken $\rinf$ finite for clarity. It is
1961: clear from \euclidgeo\ that the ${\tilde E} = 0$ geodesic is a
1962: local maximum of proper distance.
1963:
1964: At $t_E = 0$ the correct prescription to compute the correlator is
1965: to sum over the two dominant (${\tilde E} \neq 0$) geodesics. We
1966: then can continue into the Lorentz section by varying $t_L$
1967: away from zero. To compute the correlator we follow both of
1968: the local minimum geodesics. From \Eexpansthree\ we see that $E$ for this pair
1969: of geodesics will now have to have nonvanishing real and imaginary
1970: parts. If we label the two geodesics $1$ and $2$, we have
1971: $E_1=E_2^*$ and $\len_1 = \len_2^*$. The real parts of the proper
1972: distances are the same, so the magnitude of their contribution is
1973: the same. So our final formula for the two sided correlator as a
1974: function of Lorentzian time (or more generally, in a complexified
1975: neighborhood of Lorentzian time) is given by \eqn\minkcorr{\langle
1976: \phi \phi \rangle(t) = e^{-m\len_1} + e^{-m\len_2}.}
1977:
1978: Because we have resolved the geodesic bifurcations we have been
1979: able to give an unambiguous calculation of this correlator. Taking
1980: the mass of the black hole to infinity we see that the correct
1981: prescription using \longeqn\ is to sum over both complex branches
1982: with equal weight.
1983:
1984:
1985: %-------------------------------------------------------------------
1986: \subsec{A bifurcation model}
1987:
1988: As we have seen, at the branch
1989: point described in \Eexpans, the geodesics coalesce. This means
1990: that stationary points of the path integral action become
1991: arbitrarily close in path space. This indicates that an
1992: eigenvalue of the fluctuation operator around such a geodesic must
1993: go to zero. There is a soft mode in the fluctuation spectrum
1994: around the branch point geodesic. More explicitly, the soft mode
1995: corresponds to infinitesimal motions of $E$ away from zero;
1996: because $t(E) \sim E^3$ this means that the boundary points are
1997: fixed under such motions. In the space of paths, one can argue
1998: that the coordinate $E$ is a good nonsingular coordinate for this
1999: mode near zero. For example, the central point of the path is
2000: deflected by a distance proportional to $E$. Since the regularized
2001: length behaves like $\len \sim E^4$ we see that the fluctuation
2002: eigenvalue goes like $d^2 \len/ d E^2 \sim E^2 \sim t^{2/3}$. This
2003: zero mode at $t=0$ will create divergences in the $1/m$ expansion
2004: there.
2005:
2006: The fluctuation dynamics near $t=0$ should be dominated by this
2007: zero mode. So we should be able to get a reliable picture of the
2008: correlator in this region by truncating the path integral to this
2009: one fluctuating mode. Of course it will be non-gaussian, but
2010: from our knowledge of $\len(E)$ we should be able to determine its
2011: dynamics. Representing the nonsingular coordinate for the zero
2012: mode $E$ by $x = i E$, and setting $\tau=-it$ we can model the
2013: path integral for $t \sim 0$ by the ordinary integral\foot{This
2014: integral is a special case of Pearcey's Integral \BerryMII .}
2015:
2016:
2017: \eqn\f{I(\tau) = \int_{-\infty}^{\infty} e^{-m V_{\tau}(x)} dx~,}
2018: where the truncated action is given by \eqn\action{V_{\tau} =
2019: \quarter x^4 - \tau \, x.}
2020:
2021:
2022: To motivate this choice for $V_{\tau}$, compute its stationary points.
2023:
2024: \eqn\vsad{\eqalign{V_{\tau}'(x_s) &= x_s^3 - \tau =
2025: 0 \cr x_s &= \tau^{\third} \cr V(x_s) &=-{3 \over
2026: 4} \tau^\ft~.}}
2027:
2028: This is the analog of \Eexpans.\foot{ We will be able to study
2029: this integral in sufficient detail that we will not need to add
2030: the additional terms $\sim x^2$ to $V$ that would resolve the
2031: cubic branch point. This can be done though.}
2032: For $\tau$ real, ($t$ Euclidean), the integrand in \f\ is real and
2033: positive on the real $x$ integration contour. In such a
2034: situation the large $m$ saddle point analysis is straightforward.
2035: We can then study deformations of $t$ away from Euclidean.
2036:
2037:
2038: We now study this model for complex $\tau$. We can still use
2039: saddle point approximation for large $m$, but now we have to
2040: determine carefully which saddles lie on the integration contour.
2041: More precisely, for complex $t$ we must deform the integration
2042: contour so that it becomes a steepest descent path for the real
2043: part of $V_{\tau}$. Then we can evaluate $I(\tau)$ by doing a
2044: saddle point expansion around all the saddles included on the
2045: path. A steepest descent path for the real part of $V_{\tau}$ is
2046: simply a contour of constant $\Im (V_{\tau})$, so they are easy to
2047: identify.
2048:
2049: \ifig\stokesa{Integration contours for various values of $t$: left
2050: to right, $\tau = 1, e^{i \pi /8}, e^{i \pi/5}$. At $\tau = e^{i
2051: \pi /8}$, the integral crosses a Stokes line and picks up another
2052: saddle.} {\epsfxsize=13.5cm \epsfysize=3.5cm
2053: \epsfbox{stokes3.eps}}
2054:
2055: \ifig\stokesb{Positions of the Stokes and anti-Stokes (saddle
2056: exchange) lines in the complex $\tau$ plane. The dashed lines are
2057: the Stokes lines, where the contour goes from crossing one saddle
2058: to two, and therefore picks up a sub-dominant contribution. The
2059: two saddles contribute equally on the wavy lines, so that the
2060: integral there has a kink (in the large $m$ limit) as a function
2061: of $\tau$.}
2062: {\epsfxsize=8cm \epsfysize=5.8cm \epsfbox{stokes4.eps}}
2063:
2064: For convenience, we restrict ourselves to $\Im (\tau) \geq 0$, and
2065: write \eqn\definet{\tau = r \, e ^ {3 i \theta}} with $\theta \in
2066: [0, {2 \pi \over 3}], r \geq 0$. The critical points are then
2067: \eqn\criticalp{x_k = r^{\third} \, e^{i \theta + {2 \pi i k \over
2068: 3}}} where $k = 0, 1, 2$. Consider starting with real positive
2069: $\tau$, and then rotating $\tau$ counterclockwise into the complex
2070: plane. For $3 \theta < {\pi \over 8}$, the steepest descent
2071: contour contains only the saddle point $x_0$. For example, at
2072: $\theta = 0$, the steepest descent contour is simply the defining
2073: contour of the integral: the real axis. At $3 \theta = {\pi \over
2074: 8}$, however, the imaginary parts of $V_{\tau}$ become equal at
2075: $x_0$ and $x_1$, and the contour picks up the saddle point $x_1$
2076: (\stokesa). The line $3 \theta = \pi/8$---where the integral
2077: picks up a subdominant exponential in the saddle
2078: approximation---is known as a Stokes line \BerryMI. This
2079: situation persists until $3 \theta = {7 \over 8} \pi$, at which
2080: point the contour loses the point $x_0$, and only $x_1$ remains
2081: (\stokesb).
2082:
2083: By computing the real part of $V_{\tau}$, we see that the saddle
2084: at $x_0$ dominates over $x_1$ for $3 \theta < {\pi \over 2}$, and
2085: vice versa for $3 \theta > {\pi \over 2}$. The locus of this
2086: exchange of dominance is referred to as an anti-Stokes line. Right
2087: on the line both saddles contribute equally. This is the
2088: situation for the Lorentzian correlator discussed in the previous
2089: subsection.
2090:
2091: An interesting phenomenon occurs here. For ${\pi \over 4} < 3
2092: \theta < {3 \pi \over 4}$, the overall dominant saddle is $x_3$,
2093: which is not on the integration contour, and does not contribute
2094: to the integral in this approximation. This is similar to the
2095: phenomenon we encountered in the \sads\ correlator, where the
2096: dominant branch (the $t_c$ geodesic) does not contribute. That we
2097: have a situation in which a dominant saddle does not contribute to
2098: the contour integral may at first seem surprising. In fact, it
2099: is easy to see that no matter how one tries to deform the contour
2100: so as to force it to go over this dominant saddle in a meaningful
2101: direction, one will pick up a larger extremum elsewhere on the
2102: contour. This extremum is of course only an extremum for
2103: $V_{\tau}$ restricted to the contour. It is not a critical point
2104: of $V_{\tau}$ as a function of complex $\tau$. However we do not
2105: have an analogous understanding of why the $t_c$ geodesic does not
2106: contribute to the Lorentzian functional integral. There must be
2107: other large contributions that cancel it.
2108:
2109:
2110: For $m \rightarrow \infty$ subdominant saddles make no
2111: contribution to this analytic function. The true $ m \rightarrow
2112: \infty$ correlation function \foot{More precisely, we should
2113: consider the function $L(\tau)=-{1 \over m} \ln I(\tau)$ which has
2114: a smooth large $m$ limit.} is a set of analytic regions, joined
2115: together in a non-smooth way at anti-Stokes lines when saddles
2116: exchange dominance. This is somewhat analogous to a first order
2117: transition in statistical mechanics. Computing the correlation
2118: function in one region and analytically continuing ignores these
2119: boundary lines, and is somewhat like continuing into a metastable
2120: phase. This analogy is imperfect, though. The bifurcation model
2121: illustrates that the $t_c$ saddle does not give a contribution
2122: like $e^{-m\len}$ to the path integral as would a metastable
2123: phase.
2124:
2125: Now consider $I(\tau)$ for large but finite $m$. It is clear
2126: from the defining integral \f\ that $I$ is an entire function of
2127: $\tau$. The $m = \infty$ branch point at $\tau=0$ is smoothed
2128: out. The integral has an everywhere convergent power series and in
2129: a small disk $|\tau| \ll 1/m^{3 \over 4}$, $I$ is roughly
2130: constant. When $|\tau| \sim 1$ the saddle point approximation is
2131: very accurate for large $m$. Along the anti-Stokes (saddle
2132: exchange) line, the vertical axis in \stokesb , there is a rapid
2133: but smooth transition from one saddle behavior to the other. The
2134: width of this transition is order $1/m$. So there are two half
2135: plane regions where the large $m$ saddle point behavior is very
2136: accurate, a thin joining region along the anti-Stokes line and a
2137: small roughly constant disk near the origin. These features
2138: smooth out the branch cut and so prevent us from following
2139: subdominant saddles by analytic continuation.
2140:
2141: A concrete method for performing analytic continuation is by
2142: computing power series coefficients around a point and then
2143: continuing the series variable. If we expand, for example,
2144: around a point away from the anti-Stokes (saddle exchange) line at
2145: large but finite $m$ the contamination of the subdominant saddle
2146: point contribution that prevents continuation past the point of
2147: saddle exchange is of order $e^{-m}$. Near the exchange line we
2148: can model the correction more accurately as $e^{-m(1-\tau)}$ where
2149: the exchange line is at $\tau =1$. The power series expansion of
2150: this correction goes like $e^{-m}\sum (m \tau)^k/k!$ . For fixed
2151: $k$ the corrections are order $e^{-m}$ but for $k \sim m$ the
2152: corrections are order one. So the practical question about how
2153: large an $m$ is required to be effectively infinite is equivalent
2154: to how many terms of a power series are required to accurately
2155: represent the infinite $m$ function. The functions we are
2156: interested in have smooth large $m$ limits so the number of terms
2157: required is $m$ independent. This means that accurate analytic
2158: continuation can be done with results at large but finite $m$.
2159: Examples of this procedure for studying the $t_c$ singularity are
2160: discussed in Appendix F.
2161:
2162:
2163: There is an asymptotic series of $1/m$ corrections to $I(\tau)$
2164: that can be computed by Feynman diagram techniques. Because of
2165: the soft mode, each of these terms is divergent at $\tau = 0$ and
2166: has the structure $1/(m \tau^{4 \over 3})^k$. So each term lives
2167: on the same three sheeted Riemann surface as the leading answer.
2168: If we compute a given term around a dominant saddle and then
2169: analytically continue to a region where that saddle is no longer
2170: dominant we find the same answer we would get by perturbing around
2171: the subdominant saddle using a different integration contour. So
2172: the fluctuations we compute around the dominant saddle give us
2173: information about fluctuations around the subdominant one. We
2174: can clearly do the same thing with corrections in powers of $a$.
2175: We will use generalizations of this idea to argue that we can
2176: study stringy and quantum effects around the bouncing geodesic.
2177:
2178: %-------------------------------------------------------------------
2179: \subsec{The $t_c$ singularity}
2180:
2181: We now have a reliable calculation of the boundary correlator at
2182: large $m$ in the classical supergravity approximation as a
2183: function of complex boundary time $t$. The result is contained
2184: in Equation \longeqn, with additional information about which
2185: geodesics dominate. For Lorentzian time $t= t_L$ we sum over
2186: both complex solutions of \longeqn\ with equal weight. The
2187: complex branches of \longeqn\ are free of additional singularities
2188: for all finite $t_L$. But the functions $\len_i(t)$ are analytic
2189: in $t$. Given their values along any small segment of $t$ it is
2190: possible to analytically continue them over the entire Riemann
2191: surface described by \longeqn. The $t_c$ singularity can be found
2192: by such an analytic continuation from the dominant branch selected
2193: by the boundary CFT. In the discussion above of the bifurcation
2194: model we started from Euclidean $t$. Here let us give the argument
2195: starting from the Lorentzian $t$. Consider small $t$ and hence
2196: use the expansion \Eexpans, which gives us the relation $E \sim
2197: t^{\third}$. As we have argued previously, for small Lorentzian
2198: $t$ the boundary CFT selects the branch $E= e^{{2 \pi i \over 3}}
2199: |t^{\third}|$. (We suppress the other complex branch which behaves
2200: similarly.) If we denote $t=e^{i\alpha} t_L$ then we reach the
2201: branch with real $t$ and real $E$ (and hence real $\len$) by
2202: taking $\alpha =\pi$. On this branch there is a singularity at
2203: $|t| = t_c$, due to the logarithmically diverging $\len$ of the
2204: analytically continued geodesic, which ``bounces" off the
2205: singularity. This $t_c$ singularity clearly reflects aspects of
2206: the black hole singularity. So we see that information about the
2207: black hole singularity is encoded in the boundary correlators,
2208: albeit in a subtle way involving analytic continuation. Despite
2209: its subtlety, analytic continuation can be computationally
2210: effective. In Appendix F we show how a modest number of terms in a
2211: power series expansion of $\len$ on a dominant branch is enough to
2212: extract precision information about the $t_c$ singularity.
2213:
2214:
2215: %-------------------------------------------------------------------
2216: \subsec{Quasinormal modes}
2217:
2218: \ifig\timecontour{Contours of constant
2219: imaginary part for the function $t(E)$. There are three $\Im(t) = 0$ contours,
2220: which cross at the origin. The $t_c$ branch is the real axis. The other two give us
2221: the non-$t_c$ branches of the correlator. Pane b) is a magnified
2222: view of the singularity at $1+i$, showing only the $\Im(t)=0$ contour.}
2223: %{\epsfxsize=7cm \epsfysize=7cm \epsfbox{timecontour.eps}}
2224: {\epsfxsize=12cm \epsfysize=5.8cm \epsfbox{timecurl3.eps}}
2225:
2226: We can use the geodesic expression for the correlator derived
2227: above to compute the quasinormal modes of the \sads\ black hole,
2228: in the large $m$ limit. These agree with the known results
2229: \NunezEQ, providing a check of our calculation. \foot{We thank Chris Herzog
2230: for pointing out a mistake in a previous version of this section.}
2231:
2232: To compute quasinormal modes we need the large $t$ asymptotics of
2233: the correlator. We first note that along either complex branch
2234: $E(t)$ must go to one of the fourth roots of $-4$ as $t$ goes to
2235: infinity. Specifically, let $\a_1$ and $\a_2$ be $-1-i$ and $1-i$
2236: respectively; we can rewrite formula $\longeqn$ as \eqn\longeqna{t
2237: = {1 \over 4} \ln \[ (E - \a_1) (E - \bar{\a_1}) \over (E - \a_2)
2238: (E - \bar{\a_2})\] - {1 \over 4} i \ln
2239: \[ - (E - \a_1) (E - \a_2) \over (E - \bar{\a_1}) (E - \bar{\a_2})
2240: \].} We can expand around these
2241: various roots; for example, expanding $E$ around $\a_2$ gives $t =
2242: c - \( \a_2 / 4 \) \ln (E - \a_2)$, where $c$ is a complex
2243: constant. Writing $E - \a_2 = r e^{i \theta}$ and requiring that
2244: $t$ be real gives \eqn\treal{ \theta = \ln r + c',} where $c'$ is
2245: a real constant. This demonstrates that the complex branches
2246: (where $t$ goes to real infinity) are spirals in the complex $E$
2247: plane, converging on one of the roots (see \timecontour). Coupled
2248: with the expansion $\len \sim - {\half} \ln (E - \alpha)$ around
2249: any of the four roots, we have that as $t$ goes to infinity, $t
2250: \sim \half (1 + i) \len$ along one of the complex branches and $t
2251: \sim \half (1-i) \len$ along the other. This gives the leading
2252: large $m$ behavior of $e^{-m \len_1} + e^{-m \len_2} = e^{- m(1+i)
2253: t} + e^{- m(1-i) t}$. This reproduces the first large $m$ quasinormal
2254: modes of this black hole found in \NunezEQ. Note that conventions differ
2255: between \NunezEQ\ and this paper.
2256:
2257: To go beyond the first quasinormal mode, we argue as follows.
2258: Suppose we are on the complex branch that goes to $\a_2$ as $t$
2259: goes to infinity. From $\lengthEr$ we then see that $e^{-\len} =
2260: {(E - \a_2)}^{\half} \tilde{f} (E)$, where $\tilde{f}$ is
2261: analytic. We can invert and obtain \eqn\qa{E = \a_2 + e^{-2 \len}
2262: f(e^{-2\len})} with $f(0) = 1$ and $f$ analytic. Substituting, we
2263: obtain $t = {\half} (1+i) \len + \tilde{g}(e^{-2 \len})$, with
2264: $\tilde{g}$ analytic. Inverting once more we obtain \eqn\qb{\len
2265: = (1-i) t + g(e^{-2 (1-i) t}).} Now, the correlator is
2266: \eqn\qba{\langle\phi\phi\rangle(t) = e^{-m (\len(t) + (1/m) f_1
2267: (t) + \cdots)}} where $f_1$ is the first $1/m$ correction. By
2268: extensivity and locality, we know that $f_1(t) = b_0 t + h(t)$
2269: where $b_0$ is a constant of order one determined by the
2270: fluctuation determinant around the geodesic (which we have not
2271: calculated) and $h(t)$ is exponentially small in $t$ because the
2272: world line path integral is massive in \sads. Expanding gives a
2273: series
2274:
2275: \eqn\qc{\langle\phi\phi\rangle(t) \sim
2276: \sum_{n=0}^{\infty} c_n e^{-(m+b_0+2n)(1-i)t } ~~ + {\rm c.c.}}
2277: This yields values for the quasinormal modes \eqn\qom{\omega_n= (m
2278: + b_0+ 2n)(1\pm i)~~~n=0,1,2 \ldots ~.} These values agree with the
2279: large $m$ quasinormal modes calculated in \NunezEQ\ if $b_0=
2280: -1$. The results in \NunezEQ\ were obtained
2281: by a direct study of the radial equation for the \sads\ geometry, the Heun
2282: differential equation. In some way the geodesic computations
2283: summarized in \longeqn\ give a WKB solution to the Heun equation.
2284:
2285:
2286:
2287: %-------------------------------------------------------------------
2288: \newsec{Beyond the geodesic approximation}
2289:
2290: In the previous section we studied the massive scalar correlator
2291: in the limit $ m \rightarrow \infty$ where the geodesic
2292: approximation applies. We worked in the classical supergravity
2293: regime, $\alpha ' = l_s^2 = 0$ and $g_s=0$. In this section we
2294: will discuss the more general situation.
2295:
2296: \subsec{Finite mass}
2297:
2298: The scalar propagator has a path integral representation
2299: \eqn\pathpropscale{ \langle \phi(\xi_1)\phi(\xi_2) \rangle =
2300: \int_0^{\infty} dT ~\int
2301: Dx(s) \exp\left(-i m/2 \int_0^T ds ~
2302: \left( {\dot x}^2 +1 \right) \right).
2303: }
2304: Here $\xi_i$ are points on the boundary and the integral is over
2305: paths in the bulk that connect these two points. For the radially
2306: separated points discussed in the previous sections $\langle
2307: \phi(\xi_1)\phi(\xi_2) \rangle = \langle \phi \phi \rangle (t)$.
2308: The mass $m$ acts like $1/\hbar$, a saddle point parameter. Taking
2309: $m$ finite means all paths are explored in the functional
2310: integral, not just the dominant saddles. So one of the key steps
2311: of the previous section--following a saddle point into a region
2312: where it is no longer dominant--will be problematic when $m$ is
2313: finite. More explicitly, as discussed in section 6.5, the
2314: function $L(m, t) = -{1 \over m} \log \langle \phi\phi \rangle(t)
2315: $ defines a piecewise analytic function in the limit $m
2316: \rightarrow \infty$ that can be continued past the anti-Stokes
2317: line where saddle point dominance is exchanged, but at finite $m$
2318: there is no clear way to follow the subdominant saddle past this
2319: line. Operationally, the ``contamination" due to subdominant
2320: saddles must be controllably small to perform such analytic
2321: continuation.
2322:
2323: The coefficients of the $1/m$ expansion can be continued
2324: unambiguously and do contain significant physical information.
2325: From \pathpropscale\ we see that this expansion is a small
2326: fluctuation or heat kernel expansion. There are two places where
2327: divergences are expected in the coefficients of $(1/m)^k$. The
2328: first is $t=0$, where multiple geodesics coincide. As discussed in
2329: section 3.3 we expect a soft mode in the fluctuations around the
2330: geodesic here. This will cause a divergence in the $1/m$
2331: expansion. The terms in the $1/m$ expansion live on the same three
2332: sheeted Riemann surface as $\len$. As discussed above, continuing
2333: these coefficients gives the small fluctuation expansion around
2334: the analytically continued saddle.
2335:
2336: The second place we expect divergences is on the bouncing geodesic
2337: for $t \sim t_c$. Here the divergence is not due to a soft mode
2338: but to large nonlinearities present because the curvature is large
2339: near the singularity. We can expand the heat kernel
2340: \pathpropscale\ in a short time expansion \eqn\shorttime{ \langle
2341: \phi(\eta_1)\phi(\eta_2) \rangle \sim \sum_{k=0}^{\infty} c_k
2342: {\mit R}^{(k)} \left({T \over m} \right)^k, } where $\eta_i$ are
2343: bulk points near the singularity lying on the bouncing geodesic
2344: and ${\mit R}^{(k)}$ is shorthand for a curvature invariant that
2345: has scaling dimension $({\rm length})^{-2k}$. Because the only
2346: length scale near the singularity is ${\mit l}$, the shortest
2347: proper time between the geodesic and the singularity, we have
2348: ${\mit R^{(k)}} \sim {\mit l}^{-2k}$. The region of proper
2349: distance $T$ where the geodesic is near the singularity is also
2350: $\sim {\mit l}$. Putting in these estimates, \shorttime\ becomes
2351: \eqn\shorttimel{ \langle \phi(\eta_1)\phi(\eta_2) \rangle \sim
2352: \sum_{k=0}^{\infty} c_k \left( {1 \over m \, {\mit l} } \right)^k
2353: ~.} This is reasonable since $m \, {\mit l}$ is the only
2354: dimensionless combination available near the singularity. From
2355: \tdiff\ we see that $r_{min} \sim 1 / E$ and $E \sim 1/(t-t_c)$.
2356: It follows immediately from the metric that ${\mit l} \sim
2357: r_{min}^2 \sim (t-t_c)^2$. So we expect the $1/m$ expansion to
2358: have the structure \eqn\shorttimee{ \langle \phi\phi \rangle(t)
2359: \sim \sum_{k=0}^{\infty} c_k \left( {1 \over m \, (t-t_c)^2
2360: }\right)^k ~.} We should be able to compute these coefficient
2361: functions on the primary sheet and the analytically continue them
2362: to the secondary sheet to study their singular behavior. This
2363: gives an example of recovering nontrivial information about the
2364: neighborhood of the singularity from outside the horizon
2365: correlators. The information here is just the diverging curvature
2366: near the singularity, and it is clear from the above that the main
2367: contribution to the singular behavior comes from distances $\sim
2368: {\mit l}$ which can be made arbitrarily short (in supergravity
2369: approximation) by taking $t \rightarrow t_c$. It is possible that
2370: the strengthening of the singularity with increasing $k$ will have
2371: a recognizable signature in the correlator on the first sheet.
2372:
2373:
2374: %-------------------------------------------------------------------
2375: \subsec{Finite $\alpha'$}
2376:
2377: We now go beyond the supergravity approximation by letting the
2378: string length $l_s$, ($ ~\alpha' \sim l_s^2$ ) be finite. In the
2379: boundary SYM theory, this corresponds to taking the 't Hooft
2380: coupling $\lambda = g_{\rm YM}^2 N = (R/l_s)^4$ finite. We still
2381: work in the classical limit $g_s \rightarrow 0, ~ N \rightarrow
2382: \infty$.
2383:
2384: When $l_s$ is finite, the behavior of correlators in the large $m$
2385: limit depends on which large $m$ excitation we are studying. If
2386: we take a generic perturbative string state with many oscillator
2387: modes excited then the size of the excitation will diverge as $m
2388: \rightarrow \infty$ and the correlator will cease to serve as a
2389: local probe. Instead we consider D-branes, which, at $g_s =0$ are
2390: essentially pointlike, infinitely massive objects \refs{\ShenkerXQ
2391: , \KabatCU , \DouglasYP} . In particular in the $AdS_5 \times S^5$
2392: IIB string theory we study D3-branes wrapped around an $S^3$
2393: submanifold of $S^5$. Such a state is pointlike in the $AdS_5$
2394: space and realizes the BPS state carrying SO(6) charge, for large
2395: charge. These are the giant gravitons \McGreevyCW.
2396:
2397:
2398: The main effect of small but finite $l_s$ on such states is
2399: encapsulated in corrections to the supergravity Lagrangian. These
2400: include, for instance, corrections of the form $(\alpha')^k
2401: R^{(k)}$. The small parameter controlling the size of such terms
2402: is $l_s/{\mit l}$. For phenomena near the horizon ${\mit l} \sim
2403: R$ and this small parameter is of order $l_s/R \sim
2404: 1/\lambda^{1/4}$. Such terms have a number of effects on the
2405: dynamics, including small shifts of the metric and dilaton fields
2406: \GubserNZ.
2407:
2408: We can estimate the effect of such shifts on the geodesic dynamics
2409: by following the discussion in the previous section about the
2410: finite mass black hole. Just as in that situation, the most
2411: general small, smooth shift in the geodesic equations can be
2412: encapsulated by deforming \longeqn\ and \lengthEr\ into
2413: \eqn\edeform{\eqalign{t &= E^3 - a E \cr
2414: \len &= -E^4 -b E^2 }}
2415: Here $a, b$ are small parameters whose signs are undetermined. As
2416: in the discussion following \Eexpanstwo\ and \Eexpans\ this
2417: structure guarantees that the bifurcation, needed for analytic continuation,
2418: does not disappear at small but finite $l_s$. At
2419: most the cube root singularity at $t=0$ splits into two square
2420: root singularities.
2421:
2422: This persistence of the singularity at small $t$ creates a puzzle.
2423: Explicit field theory calculation of the two point correlator at
2424: small $\lambda$ (i.e., large $l_s$) shows no sign of such a
2425: singularity. So it seems there must be an $m = \infty$ phase
2426: transition at finite $\lambda$. Perhaps we have missed some
2427: effect at large $\lambda$ that removes the singularity. On the
2428: other hand it does not look difficult to construct metric and
2429: dilaton modifications that remove the singularity for a finite
2430: size deformation.
2431:
2432: The nature of $l_s$ corrections for $t$ near $t_c$ gives
2433: information about stringy behavior near the black hole
2434: singularity. As discussed above, we expect corrections to be
2435: controlled by the dimensionless ratio $l_s/{\mit l} =
2436: \lambda^{1/4}/(t-t_c)^2$. So there should be large corrections
2437: for $t \sim t_c$.
2438:
2439: Of course, this is a string tree level effect and can in principle
2440: be computed by string world sheet techniques. The D-brane
2441: trajectory is encoded in a boundary state. The analytic
2442: continuation necessary to study the real geodesic requires
2443: boundary states with complex D-brane positions. Such states have
2444: been studied recently in other contexts \LambertZR\ \GaiottoRM.
2445: Perhaps one can circumvent the analytic continuation by directly
2446: formulating the conformal field theory in Lorentzian space, and then
2447: selecting the boundary state corresponding to a D-brane following
2448: the real geodesic.
2449:
2450:
2451: %-------------------------------------------------------------------
2452: \subsec{Finite $g_s$}
2453:
2454:
2455: We now relax the last remaining constraint and consider finite
2456: $g_s$. A problem immediately arises because the masses of
2457: D-branes are $\sim 1/g_s$ and so the mass $m$ of the probe
2458: particle cannot be taken infinitely large. More precisely if the
2459: $S^3$ radius on which the $D3$-brane is wrapped is $\sim R $, then
2460: its mass $m$ is $ m R \sim (m_s R)^4/g_s \sim \lambda/g_s \sim N$.
2461: If the $S^3$ radius is string scale then $m R \sim m_s R/g_s \sim
2462: N^{1/4}$. If the SO(6) charge is $Q$ then $m R = Q$.\foot{This is
2463: only approximately true because supersymmetry is broken in the
2464: black hole background, and the BPS conditions no longer strictly apply. But the
2465: semiclassical analysis in \McGreevyCW\ demonstrates that this
2466: relation continues to be approximately true for large $Q$.} The
2467: stringy exclusion principle \MaldacenaBW\ and the giant graviton
2468: analysis reviewed above show that stability requires $Q \le N$. As
2469: discussed above, at finite $m$ the correlation function
2470: contribution from the dominant saddle is ``contaminated" by
2471: $e^{-m}$ corrections from subdominant saddles that will eventually
2472: become dominant on crossing an anti-Stokes line. We can certainly
2473: study correlators as a power series in $g_s$ since $g_s^k \gg
2474: e^{-m} \sim e^{-\lambda/g_s}$ as $g_s \rightarrow 0 $. Because the
2475: emission and absorption of virtual perturbative quanta cannot
2476: significantly alter the D-brane trajectory, we should be able to
2477: analytically continue the geodesic through the anti-Stokes line.
2478: This will allow us, in principle, to study the
2479: nature of string perturbation theory in the vicinity of the
2480: singularity by continuing each order across the anti-Stokes lines
2481: to $t \rightarrow t_c$. Again, these quantities should in
2482: principle be computable by studying higher genus string diagrams
2483: with the appropriate boundary state. Of course there will be
2484: $1/m$ corrections to these order $g_s^k$ amplitudes which can be
2485: numerically larger than higher order $g_s$ corrections. But these
2486: are fluctuations about the main saddle, and will each analytically
2487: continue to the desired real geodesic. They can be
2488: distinguished from $g_s$ effects by their different parametric
2489: dependence on $\lambda$, or equivalently $Q$.
2490:
2491: In the supergravity approximation at least, we might expect
2492: divergences in loop amplitudes when interaction vertices approach
2493: the singularity. This issue was discussed for $d=3$ in
2494: \KrausIV. On general grounds, applicable here as well, such
2495: divergences should not occur. The correlators are manifestly
2496: finite in Euclidean SYM. They are analytic and so can be singular
2497: at most on a set of complex codimension one. So a generic
2498: divergence is not possible. From the bulk point of view the
2499: potential divergences in $d=3$ are not present because the
2500: analytic continuation from Euclidean space induces a shift $r
2501: \rightarrow r + i \epsilon$ which regulates the singularity.
2502: Interaction points are integrated across the singularity into
2503: another region to implement this. Remaining imaginary parts are
2504: cancelled between past and future singularities. Although $d>3$
2505: is more complicated we think it reasonable to expect a similar
2506: mechanism here.
2507:
2508: But the elimination of divergences need not eliminate large finite
2509: contributions. Since the geodesic near $t_c$ is very close to
2510: one singularity it is natural to expect large finite contributions
2511: in quantum corrections. We think this is what is being computed
2512: by the boundary theory near $t_c$.
2513:
2514: We would like, at least in principle, to be able to obtain
2515: nonperturbative information about the singularity from the
2516: nonperturbatively well defined boundary CFT.
2517:
2518: Operationally, the size of nonperturbative phenomena whose
2519: continuation we want to study must be parametrically larger at the
2520: point we calculate than the $e^{-m}$ contamination from the
2521: subdominant saddles that would become dominant at an anti-Stokes
2522: line and prevent analytic continuation of the ``metastable" phase.
2523: For example, suppose we wish to study D-brane effects, for
2524: instance the pair production of small D-branes. Such effects are
2525: typically of size $e^{-1/g_s}$. Using the largest stable giant
2526: graviton as a probe, $e^{-m} \sim e^{-\lambda/g_s}$. Since
2527: $e^{-1/g_s}$ is parametrically larger than $ e^{-\lambda/g_s}$, it
2528: should be possible to reliably compute such processes and then
2529: analytically continue them to the real sheet. Processes that may
2530: be impossible to compute include NS brane production, expected to
2531: be of order $e^{-1/g_s^2}$. It may be possible, though, to
2532: isolate these effects by studying the boundary theory as an
2533: analytic function of $g_s$ and $\lambda$.
2534:
2535: Processes where the nonperturbative effects are dominant should be
2536: much easier to study. As a first example, consider D-instanton
2537: effects that are the leading contribution to certain anomalous
2538: processes involving the bulk axion, related to the boundary
2539: $\theta$ parameter. The effect of $\theta$ on giant graviton
2540: correlators should behave like $e^{-1/g_s}(f_0(t) + O(e^{-m}))$.
2541: Here we have normalized to the $\theta = 0$ answer. The function
2542: $f_0(t)$ can in principle be computed on the primary sheet with
2543: parametrically small error. Then it can be continued to the
2544: secondary, real sheet. If D-instanton effects are enhanced near
2545: the singularity then there should be a signal in $f_0(t)$ near
2546: $t_c$.
2547:
2548: As a second example,
2549: consider a charge $Q$ wrapped D-brane ``particle" of mass $m \,
2550: R = Q$. Take $Q \sim N$. Now imagine a process where this particle
2551: fragments into two other such particles of charges $Q_1$ and
2552: $Q_2$, with $Q_1+Q_2=Q$ and $Q_i \sim Q/2$. The black hole is not
2553: a supersymmetric background so there is no BPS condition that
2554: prevents this. But this process is certainly nonperturbative. By
2555: considering the amount of D-brane that must be created and
2556: destroyed we have provisionally estimated that the rate for this
2557: process is $\sim \exp(-Q)$.\foot{In this entire discussion we have
2558: not considered the D-brane wrapped on a sphere in ${\rm AdS_5}$
2559: discussed \refs{\GrisaruZN,\HashimotoZP}. Such a state gets light
2560: near the black hole singularity and so it should have a strong
2561: effect on the analytically continued dynamics. It may be that the
2562: nonperturbative tunneling to this configuration discussed in
2563: \refs{\GrisaruZN,\HashimotoZP} is the dominant mechanism for
2564: fragmentation near the black hole singularity. This deserves more
2565: study.} Both initial and final particles follow geodesics of the
2566: kind we have been discussing, assuming that the final boundary
2567: operators at placed at the same point. These geodesics should
2568: continue smoothly onto the secondary sheet. Schematically, we
2569: expect the amplitude for this process to look like
2570: $e^{-Q}(f_0(t)+ O (e^{-Q}))$ where we have normalized to the two
2571: particle charge $Q$ correlator. Again, $f_0(t)$ is in principle
2572: calculable on the primary sheet with parametrically small error.
2573: It can then be analytically continued to $t_c$.
2574:
2575: The kind of quantum information that can be obtained results from
2576: taking $g_s \rightarrow 0$ in a certain way. If one focuses on
2577: the behavior of amplitudes as $t \rightarrow t_c$ then perhaps the
2578: information that can be extracted about the singularity is some
2579: kind of double scaling limit that compensates for a shrinking
2580: $g_s$ by examining processes that grow large for $t \rightarrow
2581: t_c$.
2582:
2583: %-------------------------------------------------------------------
2584: \newsec{Discussion}
2585:
2586: The analysis of the preceding sections shows that a significant
2587: amount of information from behind the horizon, and in particular
2588: from the region near the singularity, is encoded into boundary CFT
2589: correlators. An initial question we must consider is how this
2590: information can emerge from the hot, thermal, apparently
2591: featureless horizon. Our understanding of this issue is limited,
2592: but it is clear that analyticity plays a central role. It is
2593: analyticity that lets us reliably follow a geodesic into a region
2594: where its contribution is exponentially subdominant. One might
2595: think that our results are merely a consequence of the analytic
2596: nature of the classical \sads\ geometry, which allows one to
2597: compute the complete metric from a small sample of it far outside
2598: the horizon. But clearly more than this is involved. We have
2599: argued in the previous section that a large amount of stringy and
2600: quantum information should be accessible by analytic continuation
2601: as well. It seems that it is the analyticity of the full quantum
2602: theory, or at least a limiting part of it, that is %at stake.
2603: the central issue.
2604:
2605:
2606: Analyticity is not a central notion in classical general
2607: relativity. For instance, in the collapse of a matter shell to
2608: form a black hole the metric is not analytic, and only one
2609: asymptotic region exists. But outside the shell the metric is
2610: just the standard black hole metric. Correlators will in general
2611: not be analytic in such a spacetime. It seems that our methods
2612: may not apply to such cases. Even with analytic initial data, the
2613: same side correlators in a collapse scenario seem to differ from
2614: those of the eternal black hole by exponentially small terms, even
2615: though the geometry behind the horizon differs markedly.
2616:
2617: However, analyticity does play a central role in quantum field theory.
2618: Multipoint correlation functions are analytic functions of the
2619: point locations. So quantities computed in AdS/CFT will
2620: generically be analytic. We can envision setting up a collapse
2621: scenario in AdS/CFT by applying a large number of boundary
2622: operators at a fixed time far in the past. We then let this
2623: matter evolve. From the boundary point of view, we excite many
2624: gluons and then let them thermalize. We have created one of the
2625: microstates counted in the black hole entropy. We then can insert
2626: two very massive particle boundary operators as probes. Their
2627: correlator will be analytic in their time separation, and should
2628: go over to the thermal ensemble in the limit of large energy,
2629: which corresponds to large black hole mass. But even away from
2630: this limit the probe correlator is precisely analytic. In
2631: principle, we can test for periodicity in imaginary time, remnants
2632: of the $t=0$ branch point and of the $t_c$ singularity. We can
2633: look for exponentially small one sided contributions that become
2634: significant when continued to opposite sides. It may well be
2635: difficult to disentangle these small terms from the $e^{-m}$
2636: contamination that obstructs analytic continuation. The
2637: different parametric dependence and ability to study different
2638: microstates might make it possible to distinguish the two. This
2639: question may be easier to study in $d=3$ where the real behind the horizon
2640: geodesic dominates for all $t$. If the effective geometry behind
2641: the horizon is modified then it should make a large effect on this
2642: two sided correlator.
2643:
2644:
2645: These questions should be related to the issue of recovering
2646: information that has fallen into the black hole. The lack of
2647: information loss in eternal black holes is signalled by
2648: Poincar{\'e} recurrences \MaldacenaKR, which occur on time scales
2649: $t_r$ exponentially longer than $t_c$, $t_r \sim \exp(N^2) t_c$.
2650: These recurrences may correspond to large fluctuations in the
2651: classical geometry \MaldacenaKR. These occur at times $\sim N^2
2652: t_c$ . It would be very interesting to connect these ideas to
2653: those discussed in this paper, although the enormous difference in
2654: time scales makes this challenging.
2655:
2656:
2657: As mentioned in Section 3, tracking the bouncing geodesic
2658: through the anti-Stokes line is a bit like following a metastable
2659: phase in a statistical mechanical system, although this analogy is
2660: imprecise. A related phenomenon has already been discussed in the
2661: AdS/CFT correspondence. Gross and Ooguri \GrossGK\ considered the
2662: expectation value of two facing Wilson loops of radius $R$ and
2663: separation $L$. In the limit of infinite string tension $\lambda
2664: \rightarrow \infty$ this quantity is determined by the minimal
2665: area string world sheet spanning the loops. As $L$ is increased
2666: this surface jumps in a discontinuous, ``first order" transition.
2667: At large $L$ the disconnected surface dominates. But one can in
2668: principle continue through this ``anti-Stokes line" and study the
2669: metastable connected surface until it pinches off into a kind of
2670: singularity. It would be interesting to consider various kinds of
2671: fluctuation corrections in this situation and see which ones can
2672: be continued through the anti-Stokes line to the pinch off point.
2673:
2674: \ifig\bat{Symmetric spacelike geodesics of \sads. The numbers
2675: indicate how many such geodesics pass through each point in that
2676: region.}{\epsfxsize=7cm \epsfysize=7cm \epsfbox{bat.eps}}
2677:
2678:
2679: We have argued that the real geodesic (which becomes almost null
2680: at $t=t_c$), contributes to boundary correlators in a subtle
2681: subleading way that can be exposed by analytic continuation to a
2682: secondary sheet. We might ask whether it ever provides the
2683: dominant contribution to a physical quantity. We are not sure of
2684: the answer to this question. But, if we temporarily put aside the
2685: well founded concerns about measuring local bulk correlation
2686: functions in quantum gravity, and just compute them in classical
2687: supergravity approximation, we can find some answers. In \bat\ we
2688: have displayed the regions for which a symmetric bulk two point
2689: function is dominated by various
2690: geodesics. Inside %the bat shaped region
2691: regions 3, 2, and the central 1,
2692: the most likely scenario is that the
2693: geodesic that dominates the correlation function lies on the real
2694: branch (real $E$) and so will evolve into the nearly null geodesic. For the
2695: dominant geodesic to get within a proper time ${\mit l} $ of the
2696: singularity, the correlator points have to be separated by a proper
2697: distance $\sim {\mit l}$. For these quantities the large stringy
2698: and quantum effects discussed in the previous section should be
2699: important factors in computing the leading result, not a subtle
2700: subleading contribution. But of course the role of such bulk
2701: quantities is a deeply confusing one.
2702:
2703:
2704: We now ask what this new information can tell us about
2705: the singularity. Unfortunately, little is known about boundary CFT
2706: correlators at strong coupling. In fact, our results in the bulk
2707: make many new predictions
2708: about them, as has often been the case in AdS/CFT. Still, it is a
2709: useful exercise to ask what could be learned if complete
2710: correlator information were available for the CFT.
2711:
2712: In previous sections we concentrated on the two point correlator.
2713: But there is no reason not to study more general correlators (an
2714: example of this was discussed in Section 4). In particular, consider
2715: the correlator with $K $ operators on the asymptotic boundary of
2716: region I and $K'$ on the boundary of region III, with a variety of
2717: $SO(6)$ charge assignments. Each operator corresponds to a large
2718: mass particle (wrapped D-brane), and can be inserted at different
2719: points on the AdS sphere. Schwarzschild time runs in opposite
2720: directions in the two asymptotic regions, so the observer in region
2721: I can prepare an in-state of K particles, send the beam through
2722: the horizon, and then the other observer can observe an out-state
2723: of $K'$ particles, assuming that we have analytically continued so
2724: the bounce geodesic dominates. If the boundary times of these
2725: operators are all near $t_c$, the geodesics are almost null and
2726: hence almost on shell. A ``meta-observer" could in principle
2727: compare the data of these two observers and compute a ``meta
2728: S-matrix." This quantity would seem to contain significant
2729: information about the fate of organized matter as it approaches
2730: the singularity.
2731:
2732: But from the point of view of a global observer using Kruskal time,
2733: these correlators correspond to particle-antiparticle annihilation
2734: near the singularity, and the correlator is a kind of vacuum
2735: persistence amplitude. Its interpretation is less clear.
2736:
2737: A question one typically asks about singularities in consistent
2738: physical theories is what resolves them. Here the black hole
2739: singularity seems deeply connected to the boundary $t_c$
2740: singularity in the analytically continued CFT amplitudes. Perhaps
2741: this $t_c$ singularity is smoothed out in the full CFT with finite
2742: $\lambda $ and $g_s$. Of course the difficulties described in the
2743: previous section in extracting finite $g_s$ information make it
2744: possible that this question can only be addressed in some double
2745: scaling limit.
2746:
2747:
2748: The definition of gauge theory correlators by analytic
2749: continuation from the Euclidean thermal theory must determine some
2750: boundary conditions for bulk fields at the black hole singularity.
2751: Some work has been done on this question in the $d=3$ case
2752: \IchinoseRG, but $d>3$ is technically more difficult. We should
2753: be able to address this issue by tying geodesic techniques to WKB
2754: approximations of the wave equation. This may be connected to
2755: the possibility of defining different black hole systems by
2756: starting with different initial states than the Hartle-Hawking
2757: state.
2758:
2759: One important tool for analyzing the boundary CFT is to go to
2760: small 't Hooft coupling $\lambda$ (at $N = \infty$) where weak
2761: coupling perturbation theory is valid. But, as mentioned in
2762: Section 4.2, at weak coupling there is no sign of the branch cut
2763: signalling the metastable phase on which we concentrate. As noted
2764: above it is not difficult to imagine $l_s$ corrections to the
2765: metric and other supergravity fields that will remove the branch
2766: point and bouncing geodesic. It would be very useful to have an
2767: understanding of this from the gauge theory side.
2768:
2769:
2770:
2771:
2772: %-------------------------------------------------------------------
2773:
2774: \vskip 1cm
2775:
2776:
2777: \centerline{\bf Acknowledgements} We would like to thank Ben
2778: Craps, Ben Freivogel, Gary Horowitz, Jared Kaplan, Per Kraus, David Kutasov,
2779: Emil Martinec, Rob Myers, Mukund Rangamani, Simon Ross,
2780: Andrei Starinets, and Lenny Susskind for helpful
2781: discussions. This work is supported in part by NSF grant
2782: PHY-9870115 and by the Stanford Institute for Theoretical Physics.
2783: Matthew Kleban is the Mellam Family Foundation Graduate Fellow.
2784:
2785:
2786:
2787: %-------------------------------------------------------------------
2788: \appendix{A}{Kruskal coordinates}
2789:
2790: We have motivated the non-square nature of the Penrose diagram by
2791: studying radial null geodesics. Since the Schwarzschild
2792: coordinates do not cover the whole spacetime,
2793: we have used complexified
2794: coordinates in order to consider the globally extended spacetime.
2795: However, for the purposes of understanding the global structure,
2796: it is more desirable to pass to real, globally-defined
2797: coordinates. The most convenient ones to use are the Kruskal
2798: coordinates, which we now present. This will allow us to see the
2799: behavior of null geodesics, without having to use complexified
2800: coordinates and to deal with coordinate singularities at the
2801: horizons.
2802:
2803: Many of the issues discussed in this and the next Appendix have previously been addressed
2804: in a general context in \KloschQV\ .
2805:
2806: Suppressing the angular directions, the metric \metric\ is
2807: \eqn\metricrt{ ds^2 = - f(r) \, dt^2 + {dr^2 \over f(r)}} with $f
2808: = r^2 - {1 \over r^2}$. Now define the Tortoise coordinate
2809: \eqn\rstar{ \rs = \int_0^{r} {dr'\over f(r')} + C} with C an
2810: integration constant we will choose later, so as to make the
2811: Kruskal coordinates everywhere real. Letting $u = t - \rs$, $v = t
2812: + \rs$, we have $ds^2 = -f \, du \, dv$. This is singular at the
2813: horizon since the determinant vanishes; but
2814: %noting that at $r \sim 1$, $f \sim f'(1) \, (r - 1) = 4 \, (r - 1)$,
2815: we can perform a further coordinate transformation $U = - e^{-2 u
2816: }$, $V = e^{2 v}$ which will cast the metric into an extendible
2817: form. In these new coordinates, with $f(r)$ and $\rs(r)$ being
2818: implicitly functions of $U$ and $V$, we have \eqn\kmetric{ ds^2 =
2819: - { f \over 4} \, e^{-4 \rs} \, dU \, dV }
2820: %= {4 \over {f'(1)^2}} \, f \, e^{-f'(1) \, \rs} (-dT^2 + dX^2)}
2821: Letting $U = T-X$ and $V = T+X$, we can write this in a more
2822: familiar form $ds^2 = g(T,X) \, (-dT^2 + dX^2)$. Note that by our
2823: choice of
2824: coordinates $U(u)$ and $V(v)$, the factor $g(T,X)$ in front of %$dU\, dV$ is
2825: $(-dT^2 + dX^2)$ is nonsingular at the horizon: although $f$ by
2826: itself vanishes at the horizon, the $e^{-4 \rs}$ factor cancels
2827: this, so that the conformal factor $g$ remains finite across the
2828: horizon. Therefore, we can extend $(U,V)$ from $U<0,V>0$, to $U,V
2829: \in (-\infty, \infty)$, subject to the constraint $0 < r <
2830: \infty$.
2831:
2832: We easily derive the transformation laws\foot{
2833: For general functions $f(r)$ in \metricrt\ with simple zeros
2834: at $r= \rh$, we would have
2835: $T^2 - X^2 = - e^{f'(\rh) \, \rs(r)}$ and
2836: $2 \tanh^{-1} {T \over X} = f'(\rh) \, t$.
2837: }
2838: between the old and the
2839: new coordinates: \eqn\transa{ - e^{4 \, \rs(r)} = T^2 - X^2}
2840: \eqn\transb{ 2 \, t = \tanh^{-1} {T \over X}} Note that $\rs$
2841: acquires an imaginary part as the integral passes through the pole
2842: at $1$. It is correspondingly convenient to define $C = {i \, \pi
2843: \over 4}$, which makes $\rs$ real outside the horizon. More
2844: specifically, we have
2845: $$ e^{4 \, \rs} = \( {r-1 \over r+ 1} \)
2846: e^{2 \, \tan^{-1} r}$$
2847:
2848: Using \transa, we can now read off the curves of constant $X^2 -
2849: T^2$ labeling a constant $r$ surface. In particular, it is easy
2850: to see that the singularity ($r=0$) corresponds to $T^2 - X^2 =
2851: 1$, while the boundary ($ r \to \infty$) of \sads\ is given by
2852: $X^2 - T^2 = e^{\pi}$. Note that each of these equations is
2853: satisfied by two disjoint curves, related to each other by $X \to
2854: -X,\ T \to -T$. The full spacetime is then bounded by the four
2855: hyperbolas in Kruskal coordinates, \eqn\STbound{ -e^{\pi} < T^2 -
2856: X^2 = U \, V < 1} where the lower bound corresponds to the
2857: asymptotic (timelike) boundary, and the upper bound to the
2858: (spacelike) singularity. This is to be contrasted with the
2859: corresponding situation for the $d=3$ BTZ spacetime, where a
2860: similar analysis would yield $-1 < T^2 - X^2 = U \, V < 1$.
2861:
2862: \ifig\figKrusk{Kruskal diagrams for \sads, with a) $d=3$ and b)
2863: $d>3$. Radial null curves lie at 45 degrees; the horizons are
2864: represented by the diagonal dashed lines. The thick (spacelike)
2865: hyperbolas correspond to the singularity, the timelike ones to the
2866: boundary. The crucial difference between three and higher
2867: dimensions is that in the former case, the hyperbolas are the same
2868: distance from the origin, which translates into the geometric
2869: property that any radial null curve bouncing around the diagram
2870: ends up at the same point it started at. In higher dimensions this
2871: doesn't happen.} {\epsfxsize=14cm \epsfysize=7.35cm
2872: \epsfbox{fig_Krusk.eps}}
2873:
2874: \figKrusk\ illustrates the contrast between the Kruskal diagram in
2875: $d=3$ and $d>3$ dimensions. For the 3 dimensional BTZ spacetime,
2876: sketched in \figKrusk a, the singularities and boundaries
2877: intersect the Kruskal coordinate axes at the same distance. On the
2878: other hand, in all higher dimensions, the Kruskal diagram is as
2879: shown in \figKrusk b, where the singularities come closer to the
2880: center of the diagram than the boundaries. As we just computed,
2881: for the $d=5$ large black hole, the ratio of these distances is
2882: $e^{\pi} > 1$.
2883:
2884: This immediately implies that there cannot be a radial null
2885: geodesic starting on the boundary at $t=0$ (which implies $T=0$)
2886: and hitting the singularity in the middle ($X=0$). A more
2887: invariant way to say this is by considering a radial null geodesic
2888: (or a sequence of geodesics), which ``bounce around\foot{
2889: Specifically, starting from any point on the right boundary,
2890: consider a future-directed ingoing null geodesic; from where this
2891: hits the future singularity, draw the past-directed outgoing null
2892: curve which eventually hits the left boundary; then the
2893: past-directed ingoing null curve which hits the past singularity;
2894: and finally the future-directed outgoing null curve which comes
2895: back to the right boundary again.} the diagram'', as shown by the
2896: thin solid diagonal lines in \figKrusk. It is easy to see that if
2897: the hyperbolas intersect at the same distance, as in \figKrusk a,
2898: then such a sequence of null curves ends up at exactly the same
2899: point where it started; otherwise, if the hypebolas intersect at
2900: different distances as in \figKrusk b, the geodesics end at a
2901: different point. Furthermore, if the boundaries are further than
2902: the singularities, these curves intersect inside the spacetime. As
2903: we will discuss in Appendix B, this simple geometrical property
2904: proves, in a more invariant way, that the Penrose diagram of
2905: \sads\ cannot be drawn as a square for $d>3$, but can for $d=3$.
2906:
2907: So far, we have presented our results only in the large black hole
2908: limit. One may wonder whether for black holes comparable to (or
2909: much smaller than) the AdS radius, this effect does not go away.
2910: In fact, as we now explain, it gets larger. The smallest value
2911: that the magnitude of the lower bound in
2912: \STbound\ can have occurs in
2913: the large mass limit, where this goes to $X^2 - T^2 = e^\pi$. For
2914: smaller black holes this value increases, and in fact gets
2915: arbitrarily large in the small mass limit: the boundary is then given
2916: by $X^2 - T^2 = e^{\pi {R \over \rh}}$ with $\rh$ and $R$ denoting
2917: the black hole and AdS radius, respectively. As we discuss next,
2918: on a Penrose diagram, such as sketched in \figPdg b, the smaller
2919: black holes would have the singularities more bowed in (using the
2920: same conformal rescaling), or equivalently the boundaries more
2921: bowed out in \figPdg c. But this is exactly what we would expect
2922: if we naively cut-off the asymptotically flat Schwarzschild
2923: Penrose diagram at larger and larger distances $R$ (effectively
2924: corresponding to smaller and smaller black holes).
2925:
2926:
2927: %------------------------------------------------------------------
2928: \appendix{B}{Penrose diagram}
2929:
2930: We have already discussed the qualitative features of the causal
2931: structure of \sads\ black holes, but let us now confirm this by
2932: finding the explicit Penrose diagrams for these spacetimes. This
2933: will be achieved by finding appropriate transformation $V \to
2934: \vv(V)$ and $U \to \uu(U)$ of the Kruskal coordinates $(V,U)$
2935: (this guarantees that radial null geodesics will be 45 degree
2936: lines), such that the boundaries of the spacetime now lie at
2937: finite coordinate distance.
2938:
2939: For example, letting $V = \tan {\vv \over 2}$ and $U = \tan {\uu
2940: \over 2}$ compactifies the spacetime into a region inside $\vv,\uu
2941: \in (-\pi,\pi)$. Letting $\vv = \tau + \rho$ and $\uu = \tau -
2942: \rho$, we can now check that the singularities $U \, V = 1$ are
2943: the straight lines $\tau = \pm {\pi \over 2}, \rho \in (- {\pi
2944: \over 2}, {\pi \over 2})$. However, the boundaries, given by
2945: $(\tan {\tau + \rho \over 2} \, \tan {\tau - \rho \over 2}) =
2946: -e^{\pi}$, in the allowed region $\vv,\uu \in (-\pi,\pi)$, can be
2947: explicitly checked to be ``bowed out''. On the other hand, if we
2948: try to straighten out the boundary, e.g.\ by taking $V = e^{\pi
2949: \over 2} \tan {\vv \over 2}$ and $U = e^{\pi \over 2} \tan {\uu
2950: \over 2}$, so that $U \, V = -e^{\pi}$ is given by $\rho = \pm
2951: {\pi \over 2}, \tau \in (- {\pi \over 2}, {\pi \over 2})$, we find
2952: that the singularity must be ``bowed in''. An exact plot
2953: (generated by the latter transformation) was presented in
2954: \figgeods b.
2955:
2956: One may ask whether one can make a more clever choice of $\uu$ and
2957: $\vv$ so as to straighten out {\it both} the singularity and the
2958: boundary. After all, the usual lore that ``curves can be
2959: straightened out by suitable conformal transformations'' should
2960: apply here as well. But in fact, we see that the present problem
2961: is sufficiently constrained so that this cannot happen.
2962:
2963: \ifig\figquad{Most general form of Penrose diagram with timelike
2964: boundaries and spacelike singularities being straight (thick lines
2965: on the diagram), since the horizons (dashed 45 degree lines) must
2966: join the opposite vertices. A null ray bouncing around (solid
2967: thin line) must come back to the same point.} {\epsfxsize=5cm
2968: \epsfysize=6cm \epsfbox{fig_quad.eps}}
2969:
2970: We can make a general geometric argument for why the singularities
2971: and the boundaries cannot be drawn as straight lines, with the
2972: whole spacetime filling some compact region on the diagram. The
2973: first ingredient is to note that the horizons must be 45 degree
2974: lines, and that the singularities and boundaries must ``meet'' at
2975: these.\foot{By definition, horizons are the boundaries of the past
2976: of infinity (so the boundary must touch the horizon on the Penrose
2977: diagram), and any causal curve which enters the black hole must
2978: end at the singularity (so similarly, the singularity must touch
2979: the horizon in a Penrose diagram); and finally, the horizons
2980: themselves cannot form part of the boundary of the spacetime, as
2981: by definition they lie entirely in the interior.} (Note that this
2982: is the reason the Penrose diagram could not be e.g.\ a rectangle,
2983: which might naively seem consistent with the observed properties
2984: of null geodesics.) The most general way the corresponding Penrose
2985: diagram with straight singularities and boundaries could look is
2986: sketched in \figquad. But one can easily show by a series of
2987: similar-triangle arguments that a null curve bouncing around this
2988: diagram (analogously to that sketched in \figKrusk) must come back
2989: to the same point. This is in contradiction to the properties of
2990: null geodesics, as we have demonstrated above, so the
2991: singularities and boundaries cannot both be straight, if the
2992: spacetime is to be mapped into a compact region.
2993:
2994: The last requirement is very important: we can easily construct
2995: maps which make both singularities and the boundaries straight,
2996: but at the cost of the ``center'' (where the two horizons
2997: intersect) being mapped to infinity. One such example is to let
2998: $U = \uu$ and $V = 1/\vv$. The diagram of the spacetime would then
2999: not be compact, and would not qualify as a Penrose diagram. In
3000: retrospect, it is clear that one can't straighten out {\it any}
3001: two curves by conformal transformation, without letting them go to
3002: infinity. For example, consider two distinct curves which have the
3003: same endpoints. Whatever conformal transformation we apply, the
3004: two curves will still have the same endpoints. But clearly, in the
3005: flat geometry there is only a single straight line between any
3006: two points, so if the conformal transformation straightens out
3007: both curves without making them go off to infinity, it must be
3008: singular (i.e.\ non-invertible).
3009:
3010: We should make one cautionary remark: we have shown that, keeping
3011: the boundaries straight, the singularities must be bowed in. The
3012: amount of bowing, however, does depend on the conformal
3013: transformation used. In other words, the crucial geometrical
3014: property we used was that radial null geodesics bouncing around
3015: the diagram end up at a different point from which they started.
3016: However, by a suitable conformal transformation, we can bring this
3017: point to appear arbitrarily close to the starting point,
3018: effectively making the Penrose diagram look almost, but not quite,
3019: like a square.
3020:
3021:
3022:
3023:
3024:
3025: %-------------------------------------------------------------------
3026: \appendix{C}{Spacelike geodesics with angular momentum}
3027:
3028: In this paper, we have been using only radial geodesics in \sads.
3029: For completeness, let us now consider geodesics to ones carrying
3030: some ``angular'' momentum.\foot{ We will use notation appropriate
3031: to finite-sized (spherically symmetric) \sads\ black hole; in the
3032: large black hole limit, the angular momentum becomes the linear
3033: momentum along one of the translationally invariant directions.}
3034: In this case, the
3035: behaviour of spacelike geodesics is altered significantly. In
3036: addition to timelike and radial components, the geodesics have an
3037: angular component, described by the constant of motion
3038: $L = \phd \, r^2$.
3039: The radial equation \spgeod\ is now modified to \eqn\spgeodL{
3040: \rd^2 = E^2 + f(r) \, \( 1 - {L^2 \over r^2} \)} which describes a
3041: 1-dimensional motion of a particle of energy $E^2$ in the
3042: effective potential \eqn\Veff{
3043: V_{\rm eff}(r) = %\( r^2 - {1 \over r^2} \) \, \({L^2 \over r^2} - 1 \)
3044: -r^2 + L^2 + {1 \over r^2} - {L^2 \over r^4}}
3045: We see that rather than being repelled by the spacelike
3046: singularity, the spacelike geodesics which penetrate sufficiently
3047: close to the singularity are now attracted, and therefore
3048: terminate at the singularity. Namely, a spacelike geodesic with
3049: energy $E$ and angular momentum $L$ starting at $r= \infty$ can
3050: propagate to $r = \ri$, where $\ri$ is now a function of both $E$
3051: and $L$, given by the solution of the
3052: %$(d+1)$-degree polynomial
3053: equation \eqn\riL{
3054: E^2 + f(\ri) \( 1- {L^2 \over \ri^2} \) = 0 }
3055: Note that for $E=0$, for any $L$, $\ri = 1$ as before (assuming
3056: $L<1)$; however for $E>0$ the behaviour is more complicated, and
3057: depends on $L$. For $L>1$, we see that $\ri > 1$ since otherwise
3058: the LHS of \riL\ would be strictly positive. This means that such
3059: geodesics can never cross the horizon, and in particular can never
3060: connect the two boundaries of \sads. Thus they are not relevant
3061: for us, and we will henceforth consider only geodesics with $L \le
3062: 1$. For $L < 1$, the geodesics must propagate past the horizon
3063: (otherwise the LHS of \riL\ would again remain positive), and
3064: there are now two possibilities: for high enough energy, the
3065: geodesic terminates at the singularity, whereas for energies below
3066: this critical energy, $E < \Ec$, it reaches only down to $\ri > L
3067: > 0$, and reemerges in the other asymptotic region. The dividing
3068: energy $\Ec$ is of course given by the maximum value of the
3069: effective potential in \Veff, $\Ec^2 = f(\ri) \, \( {L^2 \over
3070: \ri^2} -1 \).$ Note that this is consistent with our previous
3071: picture, since $\Ec \to \infty$ as $\ri \to 0$, i.e.\ as $L \to
3072: 0^+$. In the small $L$ limit, this behaves as $\Ec \sim {1 \over L
3073: \, R }$.
3074:
3075: Confining attention to $L \le 1$ and $E \le \Ec$, let us consider
3076: the time $t_0$ at which a spacelike geodesic of given $E$ and $L$
3077: has to start in order to be symmetric. By the same arguments as
3078: before, this is given by \eqn\toEL{ t_0(E,L) + i \, \beta / 4 = -
3079: \int_{\ri(E,L)}^{\infty} {E \over f(r) \, \sqrt{ E^2 + f(r) \( 1 -
3080: {L^2 \over r^2} \)} } \, dr } As before, $t_0$ vanishes for
3081: $E=0$, and becomes more negative as $E$ increases. Thus, to see
3082: the possible range of $t_0$ for various allowed values of $E$ and
3083: $L$, it suffices to consider the part of the integral around $\ri$
3084: for energies close to the saturating value $\Ec$. Since the
3085: effective potential peaks at $\Ec$, and $\ri(E \to \Ec)$
3086: approaches the $r$-value at which $V_{\rm eff}$ attains its
3087: maximum, we can approximate the effective potential near its
3088: maximum by $V_{\rm eff} = - f(r) \( 1 - {L^2 \over r^2} \) \approx
3089: \Ec^2 - \mu^2 \, (r-\ri)^2$, where the coefficient $\mu^2$ is some
3090: complicated function of $L$ and the spacetime parameters.
3091: Substituting this back into \toEL, we see that for energies $E
3092: \approx \Ec$, the integrand becomes $\Ec/f(r) \over (r-\ri)$.
3093: Since $f(r)$ is well-behaved at $r = \ri$ for finite $L$, we see
3094: that the integral \toEL\ has logarithmic divergence at $\ri$.
3095: Thus, unlike in the $L=0$ case where $t_0$ was bounded by $\tc$
3096: (cf.\ \tdiff), for $L>0$ this bound disappears, and $t_0$ can have
3097: arbitrary values.
3098:
3099: Let us now consider the behavior of the proper length along these
3100: symmetric spacelike geodesics carrying angular momentum. This is
3101: given by the generalization of \lengthE, namely \eqn\lengthEL{
3102: \len = 2 \int_{\ri}^{\rinf} {dr \over \sqrt{E^2 + f(r) \( 1 - {L^2
3103: \over r^2} \) } }} where $\rinf$ is the upper cutoff and $\ri$ is
3104: given by solution of \riL. Since we already analyzed the behavior
3105: as $E$ increases, for simplicity we will now consider the $E=0$
3106: case, to see how increasing $L$ affects the proper length $\len$.
3107: For $E=0$, $\ri = 1$ independently of $L$, so we clearly see that
3108: increasing $L$ must increase the proper length $\len$ (since the
3109: denominator of the integrand in \lengthEL\ decreases). In fact, as
3110: $L \to 1$, it is easy to see that the length diverges
3111: logarithmically at the near-horizon region.
3112:
3113: %-------------------------------------------------------------------
3114: \appendix{D}{Derivation of the geodesic time $t(E)$}
3115:
3116: In this appendix we derive the formula for the boundary time $t_0$
3117: as a function of $E$ for the case of the symmetric spacelike
3118: geodesics. We first look at the infinitely massive black hole.
3119: The integral to be done is \tstartsp ,
3120: \eqn\parta{-t_0=\int_{\ri}^{\infty} {E \over f(r) \, \sqrt{ E^2 +
3121: f(r) }} \, dr} where $\ri$ is the turning point and $f(r) = r^2 -
3122: {1 \over r^2}$. Letting $r^2 = u$, $\gamma = \sqrt {1 + \quarter
3123: E^4}$ and $u=v \, \gamma- \half E^2$, we obtain
3124: \eqn\partb{-t_0=\int_{\ri^2}^{\infty} {\half E \, u \, du \over
3125: (u^2 - 1) \sqrt {{(u + \half E^2)}^2 - (1 + \quarter E^4)}} =
3126: \int_{1}^{\infty} {\half E \, (v \, \gamma - \half E^2) \, \gamma
3127: \, dv \over ((\gamma \, v - \half E^2)^2 - 1) \, \gamma \, \sqrt
3128: {v^2 - 1}}.}
3129: Now
3130: let $v = \cosh x$. The integral becomes
3131: \eqn\partc{-t_0=\half E
3132: \int_{0}^{\infty} {(\gamma \cosh x - \half E^2) \, dx \over \gamma^2
3133: \, v^2 - \, \gamma E^2 v + \quarter E^4 - 1} = {E \over 4 \gamma}
3134: \int_0^\infty {(2v - {E^2 \over \gamma}) \, dx \over {v^2 - {E^2
3135: \over \gamma} v + \quarter E^4 - 1}}}
3136: Note that the numerator is
3137: the derivative of the denominator with respect to $v$. This makes
3138: the decomposition into partial fractions easy, and the integral
3139: becomes
3140: \eqn\partd{- t_0={E \over 4 \gamma} \int_0^\infty {dx \({1
3141: \over v - \alpha} + {1 \over v - \beta}\)}}
3142: where $\alpha = {\half
3143: E^2 + 1 \over \gamma}$ and $\beta = {\half E^2 - 1 \over \gamma}$
3144: are the two roots of the denominator of the previous equation. The
3145: integrals can now be done by substituting $w = e^x$ and writing
3146: $\cosh x$ in terms of $w$. The answer is
3147: \eqn\parte{-t_0={E \over 2
3148: \gamma} \[ {1 \over \mu_\alpha - \nu_\alpha} \ln \({1 - \nu_\alpha
3149: \over 1 - \mu_\alpha}\) + {1 \over \mu_\beta - \nu_\beta} \ln \({1
3150: - \nu_\beta \over 1 - \mu_\beta}\)\]}
3151: where $\mu_x = x + \sqrt{x^2
3152: - 1}$ and $\nu_x = x - \sqrt{x^2 - 1}$. Using the actual
3153: expressions for $\alpha$ and $\beta$ in terms of $E$, we can
3154: simplify the answer to the following form:
3155: \eqn\partf{-t_0=\quarter \ln
3156: \({\half E^2 - E + 1 \over \sqrt{1 + \quarter E^4}}\) - {i \over
3157: 4} \ln \({- \half E^2 + i E + 1 \over \sqrt{1 + \quarter E^4}}\)}
3158: Here we have subtracted the imaginary piece $i \beta/4$
3159: corresponding to crossing one horizon. One can check that this
3160: gives $t_c$ for large $E$ and that it goes like $E^3$ for small
3161: $E$. This shows that there are 3 branches corresponding to the
3162: positive $t$ direction starting at $0$ on the complex $E$ plane.
3163:
3164: Recall that the result \longeqn\ quoted in Section 3.2 pertained to
3165: the time difference between two boundaries, so that for symmetric
3166: geodesics, $t =- 2 \, t_0$ (defined to be real for real energies $E$).
3167:
3168: The case of finite mass black hole can also be done exactly, though
3169: it is computationally more involved. In particular, following
3170: the procedure outlined above for the metric \metric, we obtain
3171: \eqn\tfinite{\eqalign{ -t_0(E) = {R^2 \over 2 \Rrr} \ \{
3172: \rh \, & \ln \( {(1+E^2) \, R^2 - 2 E \, R \, \rh + 2 \rh^2 \over
3173: \sqrt{(1+E^2)^2 \, R^4 + 4 R^2 \, \rh^2 + 4 \rh^4}} \) \cr - i \,
3174: \sqrt{\rh^2 + R^2} \, & \ln \( {(1-E^2) \, R^2 + 2 \, i \, E \, R
3175: \, \sqrt{\rh^2 + R^2} + 2 \rh^2 \over \sqrt{(1+E^2)^2 \, R^4 + 4
3176: R^2 \, \rh^2 + 4 \rh^4}} \) \} }} Note that for $\bar{t} = {\rh
3177: \over R^2} t$ and $\bar{E} = {R \over \rh} E$, in the ${R \over
3178: \rh} \to 0$ limit, $-\bar{t}_0(\bar{E})$ reduces to \partf\
3179: with barred $t$ and $E$.
3180:
3181: Expanding around $E \approx \infty$ yields $t_c = -{1 \over 4 T} {
3182: \sqrt{\rh^2 + R^2} \over \rh }$ which can be easily verified by
3183: considering radial null geodesic as in \tra.
3184:
3185: Expanding around $E \approx 0$ yields \eqn\Eseries{ -t_0(E)
3186: \approx {R^5 \over \Rrr^2} \, E + {R^5 \, (2 \rh^2 - R^2)^2 \over
3187: 3 \, \Rrr^4} \, E^3 + \CO({E^5}) .}
3188: In passing we note
3189: the intriguing fact that the cubic term flips sign when $\rh = {R
3190: \over \sqrt{2}}$, which is exactly the same as the transition point when
3191: the specific heat becomes negative---that is, at the Hawking-Page
3192: transition \HawkingDH. Although the physical significance is not
3193: clear, it is rather suggestive that these small black holes (which
3194: are no longer approximately thermal states in the CFT) have a vastly
3195: richer geodesic structure.
3196:
3197: %The final result turns out to be that
3198: The important point, used in Section 3.3, is that $t(E)$ now has
3199: the form \eqn\finalresult{t(E) \sim E^3 - a E} with $a<0$, so
3200: that the cubic degeneracy is resolved for Euclidean time.
3201:
3202: %For the finite mass black hole case, we have to do some careful
3203: %rescaling of the relevant quantities, but the final result is that
3204: %\eqn\finalresult{t = aE + E^3} with $a>0$. The detailed
3205: %derivation is as follows:
3206: %
3207: %\eqn\metric{ ds^2 = - f(r) \, dt^2 + {dr^2 \over f(r)} + r^2 \,
3208: %d\Om_3^2} with \eqn\ffull{ f(r) = {r^2 \over R^2} + 1 - {\rh^2
3209: %\over r^2} \( {\rh^2 \over R^2} + 1 \)
3210: % }
3211: %where $\rh$ is the horizon radius, and $R$ is the AdS radius.
3212: %
3213: %The large mass limit consists of rescaling
3214: %$$ r = \sqrt{R \, \rh} \ \tilde{r}, \ \ \
3215: % t = \sqrt{{R^3 \over \rh}} \, \tilde{t}, \ \ \
3216: % f(r) = {\rh \over R} \, \tilde{f}(\tilde{r})$$
3217: %in addition to rescaling the energy and proper length by
3218: %$$ t = \sqrt{{\rh \over R}} \, \tilde{E}, \ \ \
3219: %\lambda = R \, \tilde{\lambda} $$ and taking the limit ${R \over
3220: %\rh} \to 0$. Under this rescaling, the geodesic equations are
3221: %written in the correct form we have been using.
3222: %
3223: %The starting time $t_0$ as a function of energy $E$ for radial
3224: %spacelike symmetric geodesics is then given by
3225: %\eqn\tfinite{\eqalign{ -t_0(E) = {R^2 \over 2 \Rrr} \ \{
3226: % \rh \, & \ln \( {(1+E^2) \, R^2 - 2 E \, R \, \rh + 2 \rh^2 \over
3227: %\sqrt{(1+E^2)^2 \, R^4 + 4 R^2 \, \rh^2 + 4 \rh^4}} \) \cr - i \,
3228: %\sqrt{\rh^2 + R^2} \, & \ln \( {(1-E^2) \, R^2 + 2 \, i \, E \, R
3229: %\, \sqrt{\rh^2 + R^2} + 2 \rh^2 \over \sqrt{(1+E^2)^2 \, R^4 + 4
3230: %R^2 \, \rh^2 + 4 \rh^4}} \) \} }} Note that for $\bar{t} = {\rh
3231: %\over R^2} t$ and $\bar{E} = {R \over \rh} E$, in the ${R \over
3232: %\rh} \to 0$ limit, $-\bar{t}_0(\bar{E})$ reduces to \longeqn.
3233: %
3234: %Expanding around $E \approx \infty$ yields
3235: %$$ -t_0(E) \approx {\pi R^2 \, \sqrt{\rh^2 + R^2} \over 2 \Rrr} +
3236: % {R \over E} + \CO({1 \over E^3}) $$
3237: %where we recognise the first term as $t_c = {1 \over 4 T} {
3238: %\sqrt{\rh^2 + R^2} \over \rh }$. This provides a check for the
3239: %answer.
3240: %
3241: %Expanding around $E \approx 0$ yields \eqn\Eseries{ -t_0(E)
3242: %\approx {R^5 \over \Rrr^2} \, E + {R^5 \, (2 \rh^2 - R^2)^2 \over
3243: %3 \, \Rrr^4} \, E^3 + \CO({E^5}) }
3244: %
3245: %Note the intriguing fact that the cubic term flips sign when $\rh = {R
3246: %\over \sqrt{2}}$, which is exactly the same transition point when
3247: %the specific heat becomes negative -- that is, the Hawking-Page transition [RIGHT?? REFS].
3248: %
3249: %Despite the linear term dominating over the cubic for small $E$ in
3250: %the finite mass case, here is how only the cubic term appears if
3251: %we expand $E(t)$: using $\bar{t} = {\rh \over R^2} t$
3252: %and $\bar{E} = {R \over \rh} E$, \Eseries\ (with few more terms)
3253: %becomes
3254: %$$ - \bar{t}_0(\bar{E}) \approx {R^2 \over 4 \rh^2} \bar{E} +
3255: %{1 \over 12} \bar{E}^3 - {3 R^2 \over 80 \rh^2} \bar{E}^5 - {1
3256: %\over 112} \bar{E}^7 + {5 R^2 \over 576 \rh^2} \bar{E}^9 + \cdots
3257: %$$ If we let ${R \over \rh}$ to strictly vanish, then only the
3258: %cubic, 7th order, etc., terms remain. However, this is a valid
3259: %approximation only for ${R^2 \over \rh^2} \ll \bar{E}^2 = \( {R
3260: %\over \rh} E \)^2$, which implies that $E \gg 1$. The point is
3261: %that $\bar{E}$ can be small even when $E$ is large, and this
3262: %allows the the cubic term to dominate in one expansion and the
3263: %linear term to dominate in the other.
3264:
3265: %-------------------------------------------------------------------
3266: \appendix{E}{Analytic structure of $E(t)$}
3267:
3268: We have given above a formula for $t(E)$ in the infinite mass
3269: black hole limit. This formula defines a Riemann surface over the
3270: $t$-plane, and it is amusing to examine its analytic structure.
3271: Note that because of formula $\lengthEr$ the analytic structure of
3272: $E(t)$ is closely related to that of the large $m$ correlation
3273: function (modulo extra branch cuts from the log in $\lengthEr$).
3274: To start, we find the locations of all the branch points of
3275: $E(t)$. These must correspond to values of $E$ for which $dt / dE$
3276: vanishes, and a quick examination of $dt / dE$ shows that it
3277: vanishes only at $E = 0, \infty$.
3278:
3279: Now $t$ is periodic with period $i \beta= i \pi$. For $E(t)$, we
3280: have first of all the coincident points pole at $\pi i / 2$ (with
3281: $E \sim {1 \over t - \pi i / 2}$); our previous analysis shows
3282: that we have a branch cut at $t=0$ (with $E \sim t^{\third}$), and
3283: the $t_c$ pole at $\pi / 2$. In fact this pattern of branch cuts
3284: and poles repeats itself, with $t^{\third}$ branch cuts
3285: corresponding to points where $E=0$ and thus (looking at the
3286: residues of the logs in $\longeqn$) occurring at $t = (p + iq) \pi
3287: / 2$ with $p + q$ even, and $1/t$ poles corresponding to points
3288: where $E = \infty$ and occurring at $t = (p + iq) \pi / 2$ with $p
3289: + q$ odd. Of course, these poles and branch cuts only occur on
3290: certain sheets of the Riemann surface. It is not too difficult to
3291: understand the sheet structure; the only hard part is seeing what
3292: happens at the branch points. There we have $3$ sheets. It turns
3293: out that two of those (the non-$t_c$ branches) do not have any
3294: singularities or branch cuts. Moving far in Lorentzian time along
3295: these branches corresponds to $E$ spiraling in to one of the
3296: fourth roots of $-4$, and thus getting far away from $E = 0,
3297: \infty$, where the poles and branch points occur. Of course, the
3298: non-$t_c$ branches are the physical ones, so in some sense all the
3299: interesting analytic structure is very nonphysical. One possible
3300: interpretation for the string of $t_c$ poles is that they
3301: correspond to geodesics bouncing off both the singularities and
3302: the boundaries of \sads\ multiple times. In order to bounce off a
3303: singularity the geodesic has to be spacelike whereas to bounce off
3304: the boundary of \sads\ it has to be timelike, so this idea is
3305: speculative. It is worth noting, however, that the poles on both
3306: boundaries predicted above are precisely the locations where the
3307: geodesic bounces.
3308:
3309: %-------------------------------------------------------------------
3310: \appendix{F}{Computing properties of the $t_c$ singularity}
3311:
3312: In the text we argued that certain information about the black
3313: hole singularity is encoded in boundary correlators. In
3314: particular there is information in a singularity at $t=t_c$ on a
3315: secondary sheet of the two point correlator, which at $m= \infty$
3316: has a third order branch point at $t=0$. In this appendix we will
3317: show that this singularity can be studied in a computationally
3318: effective way, assuming that precise numerical information about
3319: the boundary correlator as a function of Euclidean time on the
3320: primary sheet is available. Of course such information is not
3321: available (for large $\lambda$) from direct gauge theory
3322: computation at this time. But we feel it is useful to show by
3323: example how the process of analytic continuation can be made
3324: concrete given well defined gauge theory results.
3325:
3326: Specifically we assume that the a number of coefficients in the
3327: Taylor expansion of the gauge theory correlator $\langle \phi\phi
3328: \rangle(t)$ are known. $L(t)= {-1 \over m} \ln \langle \phi\phi
3329: \rangle(t)$
3330: has a smooth large $m$ limit so we consider its
3331: expansion. For convenience we expand around euclidean antipodal
3332: separation, $t=0$ even though $L(t)$ is branched there. The third
3333: order character of this branch point is known from our theoretical
3334: analysis or would be clear from numerical data. So we write our
3335: expansion in terms of $x = t^{2 \over 3}$. For Euclidean time
3336: separations the large $m$ evaluation of $L(t)$ from a real
3337: geodesic is certainly correct. So we can expand \lengthEr\ and
3338: \longeqn\ to obtain model ``numerical" gauge theory data. Stopping
3339: at order $x^{20}$ is sufficient to illustrate how things work.
3340:
3341: \eqn\modser{\eqalign{L(x)= & ~~0.693147 - 1.36284\,x^2 -
3342: 0.265333\,x^4 - 0.106447\,x^6 - 0.0454717\,x^8 \cr & -
3343: 0.0199415\,x^{10} - 0.0090257\,x^{12} -
3344: 0.00421708\,x^{14} - 0.00202004\,x^{16}\cr & -
3345: 0.000984367\,x^{18} - 0.000485493\,x^{20} + \ldots}}
3346:
3347: The analytic structure discussed in Section 3.2 and Appendix E
3348: implies that the $t_c$ singularity of $L$ on the secondary $t$
3349: sheet will appear at $x_c = t_c^{2/3} = 1.35128345.$ This
3350: singularity has the leading behavior $L(x) \sim 2 \log(x-x_c)$.
3351:
3352: To find this singularity numerically we use standard techniques.
3353: First we differentiate \modser\ so that the leading singular
3354: behavior is a simple pole, $L'(x) \sim 2/(x-x_c)$. We then fit
3355: \modser\ to a ratio of two polynomials $P$ and $Q$ of degree $M$
3356: and $N$, $L'(x)\simeq
3357: P_M(x)/Q_N(x)$. This is called a Pad{\'e}
3358: approximant \Baker\ . The convergence of such approximants is
3359: nonuniform, so we select one more or less arbitrarily.
3360:
3361: The $[M,N]=[7,12]$ Pad{\'e} has a pole at $x=1.35128349$ (compared
3362: to $x_c= 1.35128345$) with a residue of $2.0028$ (compared to $2$)
3363: . These results are probably anomalously good. The $[9,10]$
3364: approximant has a pole value of $1.35137$.
3365:
3366: So in this limiting case a modest amount of ``numerical" data is
3367: sufficient to study the singularity. But it is important to
3368: ascertain how stable this situation is. In particular, as
3369: discussed in Section 4.3, finite $g_s$ limits our ability to take
3370: $m$ arbitrarily large. We know that at finite $m$ the branch cut
3371: at $t=0$ is erased and we can no longer analytically continue to a
3372: secondary sheet. As discussed in Section 3.5, a signal of this
3373: erasure order one corrections to the order $k$ power series
3374: coefficient when $k \sim m$. Here we see that $k=20$ is
3375: adequate to determine $L$, so we just need $m>>20$. As also noted
3376: in section 3.5 there will be an asymptotic series of $1/m$
3377: corrections that are branched and can be analytically continued to
3378: the $t_c$ singularity. These contain interesting physics, as
3379: explained in Section 4. However to study the leading behavior
3380: they must be handled in some way, perhaps by extrapolation in
3381: $1/m$. But this may not be done with complete precision so we
3382: investigate our ability to still see the $t_c$ singularity with
3383: some small $1/m$ correction remaining. In the $x$ variable this
3384: correction has a pole at $x=0$ and a higher order pole at $x=x_c$.
3385: The correct procedure would be to expand around a point $t_0 \neq
3386: 0$ and Pad{\'e} approximate both $x=0$ and $x=x_c$ singularities.
3387:
3388:
3389: Rather than go through this entire procedure we model the
3390: phenomenon by adding the series expansion of a pole at $x={1 \over
3391: 2} $,~ $1 \over m(x-1/2)$ to the derivative of \modser. This pole
3392: is much closer to the expansion point $x=0$ than $x_c$ is so its
3393: power series coefficients grow much faster than those in \modser.
3394: The ratio of the coefficients at order $n$ is approximately ${1
3395: \over m} (x_c/(1/2))^n ~ \sim {1 \over m} (2.7)^n $ At $n=19$ for
3396: $m=100$ the $1/m$ correction is about $10^7$ times larger than the
3397: main result! Nonetheless the Pad{\'e} approximant can efficiently
3398: separate out these different poles. The $[7,12]$ approximant
3399: (with $m=100$) yields a pole at $x=1.35169$ with residue $2.0086$,
3400: as well as the pole at $x=1/2$.
3401:
3402: As explained in Section 4.1 we expect that the high curvature
3403: around the black hole singularity will create a leading $1/m$
3404: correction of form ${1/(m(t-t_c)^2)}$. The series expansion of
3405: this correction will eventually dominate over the terms in
3406: \modser. The ratio of terms of order $n$ will be roughly ${1
3407: \over m} n^2$. There exist techniques to disentangle such
3408: confluent singularities in a Pad{\'e} approximant \Baker\ . But
3409: we will content ourselves with showing that this modest rate of
3410: growth allows us to ignore $1/m$ corrections for $m \sim 10^4$.
3411: The $[7,12]$ approximant with this addition gives a pole at
3412: $x=1.35061$ with residue $1.990$. This analysis illustrates how
3413: the apparently rather abstract notions of the $ m \rightarrow
3414: \infty$ limit and subsequent analytic continuation to a
3415: singularity on a secondary sheet can be made computationally
3416: explicit. Of course the input data, Euclidean gauge
3417: theory correlators, do not yet exist!
3418:
3419:
3420:
3421:
3422: \listrefs
3423:
3424: \end
3425: