hep-th0702221/d.tex
1:  \documentclass[a4paper,12pt]{article}
2: \title{Black Hole / String Transition and Rolling D-brane}
3: \author{Yu Nakayama}
4: \pagestyle{plain}
5: \def\slash#1{\ooalign{\hfil/\hfil\crcr$#1$}}
6: \usepackage{amsmath}
7: \usepackage{amssymb}
8: \usepackage{graphicx}
9: \usepackage{hyperref}
10: %\usepackage{showkeys}
11: \setlength{\oddsidemargin}{0pt}
12: \setlength{\evensidemargin}{0pt}
13: \setlength{\topmargin}{10pt}
14: \setlength{\headheight}{0pt}
15: \setlength{\headsep}{0pt}
16: \setlength{\footskip}{30pt}
17: \setlength{\textheight}{650pt}
18: \setlength{\textwidth}{470pt}
19: 
20: 
21: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
22: % Young tableaux
23: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
24: %  draw box of size #1pt and line thickness #2pt
25: 
26: \def\drawbox#1#2{\hrule height#2pt
27:         \hbox{\vrule width#2pt height#1pt \kern#1pt
28:               \vrule width#2pt}
29:               \hrule height#2pt}
30: 
31: 
32: \def\Fund#1#2{\vcenter{\vbox{\drawbox{#1}{#2}}}}
33: \def\Asym#1#2{\vcenter{\vbox{\drawbox{#1}{#2}
34:               \kern-#2pt       % line up boxes
35:               \drawbox{#1}{#2}}}}
36: \def\Sym#1#2{\vcenter{\mbox{\drawbox{#1}{#2}
37:                          \drawbox{#1}{#2}}}}
38: 
39: \def\funda{\Fund{6.5}{0.4}}
40: \def\asymm{\Asym{6.5}{0.4}}
41: \def\bfunda{\overline{\fund}}
42: \def\basymm{\overline{\asym}}
43: \def\symm{\funda\kern-0.4pt\funda}
44: 
45: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
46: %%%   NewCommand   %%%%%%%%%%%%%%%%%%%%%%
47: 
48: \newcommand{\Om}{\Omega}
49: \newcommand{\om}{\omega}
50: \newcommand{\al}{\alpha}
51: \newcommand{\ep}{\epsilon}
52: \newcommand{\de}{\delta}
53: \newcommand{\la}{\lambda}
54: \newcommand{\La}{\Lambda}
55: 
56: 
57: \newcommand{\deebar}{\bar{\partial}}
58: 
59: \newcommand{\df}{\stackrel{\rm def}{=}}
60: \newcommand{\co}{{\scriptstyle \circ}}
61: 
62: \newcommand{\lb}{\lbrack}
63: \newcommand{\rb}{\rbrack}
64: \newcommand{\rn}[1]{\romannumeral #1}
65: \newcommand{\msc}[1]{\mbox{\scriptsize #1}}
66: \newcommand{\dsp}{\displaystyle}
67: \newcommand{\scs}[1]{{\scriptstyle #1}}
68: 
69: \newcommand{\bc}{\mathbb{C}}
70: \newcommand{\br}{\mathbb{R}}
71: \newcommand{\bz}{\mathbb{Z}}
72: \newcommand{\bq}{\mbox{{\bf Q}}}
73: \newcommand{\bn}{\mbox{{\bf N}}}
74: %\newcommand{\bsz}{\msc{{\bf Z}}}
75: \newcommand{\bsz}{\mathbb{Z}}
76: \newcommand{\bsr}{\msc{{\bf R}}}
77: \newcommand{\bsc}{\msc{{\bf C}}}
78: \newcommand{\ba}{\mbox{{\bf a}}}
79: \newcommand{\bsa}{\msc{{\bf a}}}
80: \newcommand{\bm}{\mbox{{\bf m}}}
81: \newcommand{\bsm}{\msc{{\bf m}}}
82: \newcommand{\da}{\dot{a}}
83: \newcommand{\dal}{\dot{\alpha}}
84: \newcommand{\db}{\dot{b}}
85: \newcommand{\dbeta}{\dot{\beta}}
86: 
87: 
88: \newcommand{\cA}{{\cal A}}
89: \newcommand{\cL}{{\cal L}}
90: \newcommand{\cG}{{\cal G}}
91: \newcommand{\cJ}{{\cal J}}
92: \newcommand{\cT}{{\cal T}}
93: \newcommand{\cO}{{\cal O}}
94: \newcommand{\cN}{{\cal N}}
95: \newcommand{\cM}{{\cal M}}
96: \newcommand{\cF}{{\cal F}}
97: \newcommand{\cP}{{\cal P}}
98: \newcommand{\cS}{{\cal S}}
99: \newcommand{\cR}{{\cal R}}
100: \newcommand{\cC}{{\cal C}}
101: \newcommand{\cQ}{{\cal Q}}
102: \newcommand{\cD}{{\cal D}}
103: \newcommand{\cE}{{\cal E}}
104: \newcommand{\cH}{{\cal H}}
105: \newcommand{\cU}{{\cal U}}
106: \newcommand{\cZ}{{\cal Z}}
107: \newcommand{\cI}{{\cal I}}
108: \newcommand{\cK}{{\cal K}}
109: 
110: \newcommand{\tL}{\tilde{L}}
111: \newcommand{\tJ}{\tilde{J}}
112: \newcommand{\tI}{\tilde{I}}
113: \newcommand{\tG}{\tilde{G}}
114: \newcommand{\tF}{\tilde{F}}
115: \newcommand{\tN}{\tilde{N}}
116: \newcommand{\tU}{\tilde{U}}
117: \newcommand{\tV}{\tilde{V}}
118: \newcommand{\tK}{\tilde{K}}
119: \newcommand{\tY}{\tilde{Y}}
120: \newcommand{\tQ}{\tilde{Q}}
121: 
122: \newcommand{\tq}{\widetilde{q}}
123: \newcommand{\tj}{\widetilde{j}}
124: \newcommand{\tm}{\widetilde{m}}
125: \newcommand{\tl}{\widetilde{l}}
126: \newcommand{\tchi}{\widetilde{\chi}}
127: 
128: 
129: \newcommand{\hC}{\hat{\cal C}}
130: \newcommand{\hD}{\hat{\cal D}}
131: \newcommand{\hE}{\hat{\cal E}}
132: 
133: 
134: \newcommand{\hU}{\widehat{U}}
135: 
136: 
137: \newcommand{\dpsi}{\psi^{\dag}}
138: 
139: \newcommand{\Ttop}{T^{\msc{top}}}
140: \newcommand{\Ttot}{T^{\msc{tot}}}
141: \newcommand{\Gtot}{G_{\msc{tot}}}
142: 
143: \newcommand{\ket}[1]{{|#1\rangle}}
144: \newcommand{\bra}[1]{{\langle#1|}}
145: \newcommand{\dket}[1]{{\left.\left|#1\right\rangle\right\rangle}}
146: \newcommand{\dbra}[1]{{\left\langle\left\langle#1\right|\right.}}
147: 
148: \newcommand{\dketm}[1]{{\left.\left|#1\right\rangle\right
149:  \rangle_{\msc{\bf M}}}}
150: \newcommand{\dketd}[1]{{\left.\left|#1\right\rangle\right
151:  \rangle_{\msc{\bf d}}}}
152: \newcommand{\dketc}[1]{{\left.\left|#1\right\rangle\right
153:  \rangle_{\msc{\bf c}}}}
154: \newcommand{\dbram}[1]{{{}_{\msc{\bf M}}
155: \left\langle\left\langle#1\right|\right.}}
156: \newcommand{\dbrad}[1]{{{}_{\msc{\bf d}}
157: \left\langle\left\langle#1\right|\right.}}
158: \newcommand{\dbrac}[1]{{{}_{\msc{\bf c}}
159: \left\langle\left\langle#1\right|\right.}}
160: \newcommand{\ketA}[1]{{|#1\rangle}_{A}}
161: \newcommand{\braA}[1]{{}_{A}{\langle#1|}}
162: \newcommand{\ketB}[1]{{|#1\rangle}_{B}}
163: \newcommand{\braB}[1]{{}_{B}{\langle#1|}}
164: 
165: \newcommand{\kett}{\ket{0}_{\msc{top}}}
166: \newcommand{\brat}{{}_{\msc{top}}\bra{0}}
167: \newcommand{\kettA}{\ket{0}_{A, \,\msc{top}}}
168: \newcommand{\bratA}{{}_{\msc{top}\, A}\bra{0}}
169: \newcommand{\kettB}{\ket{0}_{B, \,\msc{top}}}
170: \newcommand{\bratB}{{}_{\msc{top}\, B}\bra{0}}
171: 
172: 
173: \newcommand{\dketA}[1]{{\left.\left|#1\right\rangle\right
174:  \rangle_{A}}}
175: \newcommand{\dketB}[1]{{\left.\left|#1\right\rangle\right
176:  \rangle_{B}}}
177: \newcommand{\dbraA}[1]{{{}_{A}
178: \left\langle\left\langle#1\right|\right.}}
179: \newcommand{\dbraB}[1]{{{}_{B}
180: \left\langle\left\langle#1\right|\right.}}
181: 
182: \newcommand{\chm}[1]{\mbox{ch}^{#1}_{\msc{\bf M}}}
183: \newcommand{\chg}[1]{\mbox{ch}^{#1}_{\msc{\bf G}}}
184: \newcommand{\chim}{\chi_{\msc{\bf M}}}
185: \newcommand{\chig}{\chi_{\msc{\bf G}}}
186: \newcommand{\chic}{{\chi_{\msc{\bf c}}}}
187: \newcommand{\chid}{{\chi_{\msc{\bf d}}}}
188: 
189: \newcommand{\Chm}[1]{\mbox{Ch}^{#1}_{\msc{\bf M}}}
190: \newcommand{\Chg}[1]{\mbox{Ch}^{#1}_{\msc{\bf G}}}
191: 
192: 
193: \newcommand{\Th}[2]{\Theta_{#1,#2}}
194: \renewcommand{\th}{{\theta}}
195: \newcommand{\tTh}[2]{\widetilde{\Theta}_{#1,#2}}
196: \newcommand{\ch}[2]{\mbox{ch}^{#1}_{#2}}
197: \newcommand{\tch}[2]{\widetilde{\mbox{ch}}^{#1}_{#2}}
198: \newcommand{\tr}{\mbox{Tr}}
199: %\newcommand{\mod}{\mbox{mod}}
200: 
201: 
202: \renewcommand{\Im}{\mbox{Im}}
203: \newcommand{\sIm}{\msc{Im}}
204: \renewcommand{\Re}{\mbox{Re}}
205: \newcommand{\sRe}{\msc{Re}}
206: \newcommand{\nn}{\nonumber\\}
207: \newcommand{\NS}{\mbox{NS}}
208: \newcommand{\tNS}{\widetilde{\mbox{NS}}}
209: \newcommand{\R}{\mbox{R}}
210: \newcommand{\tR}{\widetilde{\mbox{R}}}
211: \newcommand{\sNS}{\msc{NS}}
212: \newcommand{\stNS}{\widetilde{\msc{NS}}}
213: \newcommand{\sR}{\msc{R}}
214: \newcommand{\stR}{\widetilde{\msc{R}}}
215: \newcommand{\Ad}{\mbox{Ad}}
216: \newcommand{\sdprod}
217:   {\hbox{$\hspace{1mm}\rule{0.15mm}{2mm}\hspace{-0.7mm} \times$}}
218: \newcommand {\eqn}[1]{(\ref{#1})}
219: 
220: 
221: \newcommand{\be}{\begin{align}}
222: \newcommand{\ee}{\end{align}}
223: %\newcommand{\nn}{\nonumber\\}
224: %\newcommand {\eqn}[1]{(\ref{#1})}
225: \newcommand {\rmd} {{\rm d}}
226: \newcommand{\bea}{\begin{align}}
227: \newcommand{\eea}{\end{align}}
228: \newcommand{\f}{\frac}
229: \newcommand{\dd}{\mbox{d}}
230: 
231: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
232: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
233: \allowdisplaybreaks[0]
234: 
235: 
236: 
237: \newcommand{\sectiono}[1]{\section{#1}\setcounter{equation}{0}}
238: \renewcommand{\theequation}{\thesection.\arabic{equation}}
239: \def\Fus#1#2#3#4#5#6{
240: F_{#5#6}\left[
241: \begin{array}
242: [c]{cc}%
243: #3 & #2\\
244: #4 & #1%
245: \end{array}
246: \right]}
247: %%%%%%    TEXT START    %%%%%%
248: \begin{document}
249: 
250: \begin{titlepage}
251: \thispagestyle{empty}
252: \begin{flushright}
253: UT-07-09\\
254: hep-th/0702221\\
255: %January, 2004 
256: \end{flushright}
257: 
258: \vskip 1.5 cm
259: \vspace*{2.5cm}
260: \begin{center}
261: \noindent{\textbf{\LARGE{ Black Hole - String Transition  \\\vspace{0.5cm}
262:  and Rolling D-brane
263: \vspace{0.5cm}\\
264: }}} 
265: \vskip 1.5cm
266: \noindent{\large{Yu Nakayama}\footnote{E-mail: nakayama@hep-th.phys.s.u-tokyo.ac.jp}}\\ 
267: \vspace{1cm}
268: \noindent{\small{\textit{Department of Physics, Faculty of Science, University of 
269: Tokyo}} \\ \vspace{2mm}
270: \small{\textit{Hongo 7-3-1, Bunkyo-ku, Tokyo 113-0033, Japan}}}
271: \end{center}
272: %\maketitle
273: \vspace{1cm}
274: \newpage
275: 
276: \vspace*{4cm}
277: \begin{abstract}
278: We investigate the black hole - string transition in the two-dimensional Lorentzian black hole system from the exact boundary states that describe the rolling D-brane falling down into the two-dimensional black hole. The black hole - string phase transition is one of the fundamental properties of the non-supersymmetric black holes in  string theory, and we will reveal the nature of the phase transition from the exactly solvable world-sheet conformal field theory viewpoint. Since the two-dimensional Lorentzian black hole system ($SL(2;\br)_k/U(1)$ coset model at level $k$) typically appears as near-horizon geometries of various singularities such as NS5-branes in string theory, our results can be regarded as the probe of such singularities from the non-supersymmetric probe rolling D-brane.
279: The exact construction of boundary states for the rolling D0-brane falling down into the two-dimensional D-brane enables us to probe the phase transition at $k=1$ directly in the physical amplitudes. During the study, we uncover three fundamental questions in string theory as a consistent theory of quantum gravity: small charge limit v.s. large charge limit of non-supersymmetric quantum black holes, analyticity v.s. non-analyticity in physical amplitudes and physical observables, and unitarity v.s. open closed duality in time-dependent string backgrounds. This work is based on the PhD thesis submitted to Department of Physics, Faculty of Science, University of Tokyo, which was defended on January 2007. 
280: 
281: \end{abstract}
282: 
283: \end{titlepage}
284: 
285: \tableofcontents
286: \newpage
287: 
288: \sectiono{Introduction}\label{sec:1}
289: {\bf From the Heaven}
290: 
291: {\it A luminous star, of the same density as the Earth, and whose diameter should be two hundred and fifty times larger than that of the Sun, would not, in consequence of its attraction, allow any of its rays to arrive at us; it is therefore possible that the largest luminous bodies in the universe may, through this cause, be invisible} (Laplace: 1798). It was Laplace who first predicted the existence of the black hole from the Newtonian mechanics. More than a hundred years later, in 1915 when he was serving in Russia for World War I, Schwarzshild discovered the exact static black hole solution in Einstein's general relativity. Ever since, the black hole has continued to attract  our broad attention in theoretical physics.
292: 
293: Black holes are fascinating and indeed mysterious. It is remarkable that some properties of the black hole are quite reminiscent of those of the thermodynamics: it has a definite temperature, energy and entropy, and moreover it satisfies the thermodynamical laws. To understand this coincidence, it had been long suggested that the quantum gravity would explain the microscopic statistical origins of the thermodynamic properties of the black hole.
294: Furthermore black holes challenge the validity of the quantum mechanics. The Hawking radiation, predicted from the quantum mechanics, leads to evaporation of the black hole, which ironically results in the failure of the unitary evolution of the quantum system. These mysterious natures of the black holes have continued to enchant generations of theoretical physicists.
295: 
296: Over this past two decades, theoretical physicists have gained more and more confidence in string theory as a candidate for the final  theory of everything. The theory of everything, from its tacit implication, should include a consistent theory of quantum gravity with sufficient predictive power. The best arena to test the quantum gravity is quantum black hole systems, where the semiclassical analysis leads to the puzzling issues raised above. Whether the string theory resolves these issues or not is a big challenge to string theorists.
297: 
298: One of the greatest achievements of the string theory so far is to yield a microscopic explanation of the entropy for (near) BPS black holes with large charges. The string theory, along with various dualities, has enabled us to ``count" microscopic states forming such black holes. The counting successfully agrees with the classical Bekenstein Hawking entropy formula of the corresponding macroscopic black hole. 
299: 
300: The situations, however, still remain unclear when one studies non-BPS black holes with small charge. The large quantum corrections, both in string coupling constant and large curvature effects, prevent us from the quantitative enumeration of quantum states corresponding to the black hole. Qualitatively, it has been suggested that the so-called black hole - string phase transition occurs when we consider such a small charge black hole. One of the motivation of this thesis is to understand the black hole - string phase transition in exactly solvable string theory backgrounds. 
301: 
302: 
303: In this thesis, we study the exact dynamics of rolling D-brane in the two-dimensional black hole system. The two-dimensional black hole system not only gives a toy model for the exactly solvable black hole systems in string theory, but also it can be embedded in the full superstring theory as a solution corresponding to black NS5-branes. Although our model is rather specific, we will uncover many important and universal features of quantum gravity such as the black hole - string transition. In particular We would like ask three fundamental questions about the nature  of the quantum gravity, or string theory as a candidate for the theory of everything. 
304: 
305: 
306: The first problem we would like to ask in this thesis is the small charge limit of the non-supersymmetric black hole and its relation to the black hole - string transition.
307: By studying the black hole - string transition in the two-dimensional black hole, we would like to explicitly show the phase transition between the large charge limit and the small charge limit of the non-BPS black hole systems. The origin of the phase transition is the existence of two characteristic temperatures in the string theory: the one is the Hawking temperature associated with the Hawking radiation from the black hole, and the other is the Hagedorn temperature of the underlying string theory. The relation between the two temperatures is of utmost importance in understanding the black hole - string phase transition, and we will show that the phase transition occurs exactly when these two temperatures coincide in the two-dimensional black hole system by examining the properties of the exact probe rolling D-brane boundary states. 
308: 
309: A related issue is whether the genuine two-dimensional non-critical string theory (i.e. the target space is two-dimensional) admits a black hole solution. The question has remained long unanswered. Actually, the two-dimensional black hole in the two-dimensional non-critical string theory is located well below the black hole - string phase transition point, suggesting the difficulty of physical interpretations as a black hole. Our study will also support this argument in a negative way.
310: 
311: The second problem we would like to investigate in this thesis is the relation between the analyticity and non-analyticity in amplitudes and physical quantities. It is well-known that in the supersymmetric situations, holomorphy (analyticity) plays a crucial role in determining exact BPS properties of the theory. On the other hand, to discuss phase transitions such as the black hole - string transition, the non-analyticity of the physical quantities is essential. Throughout this thesis, the interplay between the analyticity and non-analyticity appears intermittently. Especially, we highlight the universality of the decaying D-brane and the subtleties associated with Wick rotation in curved spaces in this context.
312:  
313: The third problem we would like to study is the consistency between the unitarity and the open-closed duality. The unitarity is one of the crucial ingredients of the quantum theory. In the first quantized string theory, however, the unitarity in time-dependent background is not always manifest, especially in the Euclidean world-sheet formulation.
314: The simplest consequence of the unitarity is the optical theorem. In the time-dependent physics associated with the D-brane decay, however, it is not apparently obvious whether the analytic continuation involved is consistent with the requirement from the unitarity. Indeed, the abuse of the careless Wick rotation between the Lorentzian world-sheet theory and the Euclidean world-sheet theory, would result in inconsistent results, violating the optical theorem, which will be only fixed after the careful studies of the neglected pole contributions that appear through the process of Wick rotation. The rolling D-brane in the two-dimensional black hole system is an excellent arena to check the validities of proposed prescription for the Wick rotations given in the literatures.
315: 
316: {\bf Down to Earth}
317: 
318: 
319: So far, we have stated the celestial motivations of the thesis. What about the terrestrial motivations? In other words, which practical physics can we learn from the study of the rolling D-brane in two-dimensional black holes?
320: 
321: The dynamics of the rolling D-brane in the two-dimensional black holes closely resembles that of the rolling tachyon associated with the D-brane decay in flat space. Indeed, our study suggests that this tachyon - radion correspondence shows rather universal features of closed string radiation rate from the decaying D-brane. The string (particle) production from the time-dependent system such as the dynamical D-brane system itself is an interesting arena of theoretical physics, but it also has potential applications to the quantum cosmology based on the superstring theory.
322: 
323: In the recent observational cosmology, the existence of the inflational epoch of our universe has been confirmed with increasingly great accuracy. It is, therefore, a great challenge for the string theory to provide a natural setup for the inflation. One viable scenario for the string inflation is the so-called brane inflation, where the potential between the D-brane and anti D-brane provides the inflaton field. Recent studies show that the brane inflation could be embedded in the flux compactification of the type II string theory with all moduli fixed.
324: 
325: The end-point of the brane inflation is the pair annihilation between the D-brane and the anti D-brane. This is the point where the effective field theory approximation for the brane inflation breaks down and the stringy effects dominate. The reheating of the universe associated with the inflation decay is astonishingly different in the brane inflation scenario from the conventional field theory scenario. To understand the reheating process with the open string tachyon condensation, the universality of the radiation rate of the D-brane decay we will discuss in this thesis is crucial.
326: 
327: We will also see that large closed string loops will form during the D-brane decay and they will dominate the radiated energy once the fundamental string charge is induced. The subsequent evolution of such macroscopic strings will be of great importance to understand and estimate the relic cosmic strings in our universe, which might be observed in near future by experiment, directly proving the string theory.
328: 
329: In this way, the study of the D-brane decay has potential applications to quantum cosmology. We believe that our results, especially the universal properties of the decaying D-branes will become fundamental backgrounds for the realistic brane inflation models with successful reheating.
330: 
331: {\bf Organization of the Thesis}
332: 
333: 
334: The organization of this paper is as follows. In section \ref{sec:2}, we review the two-dimensional black hole from the space-time viewpoint. In section \ref{sec:3}, we review the two-dimensional black hole from the conformal field theory viewpoint. In section \ref{sec:4}, we introduce the concept of the black hole - string transition. In section \ref{sec:5}, we study the rolling tachyon dynamics and introduce the tachyon - radion correspondence conjecture. In section \ref{sec:6}, we study the D-branes in two-dimensional black hole system in the Euclidean signature. In section \ref{sec:7}, we construct the exact boundary states for the rolling D-brane in two-dimensional black hole in the Lorentzian signature. In section \ref{sec:8}, we study the closed string radiation rate from the rolling D-brane and probe the black hole - string transition. In section \ref{sec:9}, we present some discussions and the conclusion of the paper.
335: 
336: In appendices we collect useful facts used in the main part of the thesis. In appendix \ref{sec:A}, we fix our conventions and collect useful formulae. In appendix \ref{sec:B}, we present miscellaneous topics, whose detailed discussions are omitted in the main stream of the thesis.
337: 
338: A part of the thesis is based on the published papers. In particular, a large portion of the discussions in section \ref{sec:7} and \ref{sec:8} is based on \cite{Nakayama:2005pk,Nakayama:2006qm}.
339: 
340: \newpage
341: \sectiono{Two-dimensional Black Hole: Space-Time Viewpoint}\label{sec:2}
342: In this section, we review the two-dimensional black hole from the space-time viewpoint. We will see that the string theory is replete with exactly solvable solutions containing the two-dimensional black hole systems. By studying such backgrounds, we can understand the $\alpha'$ exact physics of the string theory near singularities.
343: 
344: The organization of this section is as follows. In section \ref{sec:2-1}, we introduce the black NS5-brane background as a most typical string solution based on the two-dimensional black hole system. In section \ref{sec:2-2}, we generalize the construction to study string theory near various singularities. In section \ref{sec:2-3}, we review the basic aspect of the classical two-dimensional black hole system. In particular we focus on the thermodynamic properties in section \ref{sec:2-4}.
345: \subsection{(Black) NS5-brane background}\label{sec:2-1}
346: As is often said, the string theory is not a theory of strings only. It turns out to contain other higher dimensional nonperturbative objects such as D-branes and NS-branes. Stable D-branes are charged under the Ramond-Ramond fields, and defined as objects on which perturbative strings can end. On the other hand, NS-branes are charged under the Kalb-Ramond $B_{\mu\nu}$ field, and do not possess a perturbative definition. They can be constructed as solitonic solutions of the equation of motions of the effective supergravity in ten-dimension.
347: 
348: Historically, all these important ingredients of the string theory are discovered as exact (BPS) solitonic solutions of the effective supergravity in ten-dimension. The tension of the D-branes is proportional to $1/g_s$ while the tension of the NS-branes is proportional to $1/g_s^2$, where $g_s$ denotes the string coupling constant. Hence, in the perturbative limit (i.e. $g_s \to 0$), all these objects are infinitely massive compared with the perturbative string spectrum and could  be neglected as excitations. Rather we regard the existence of such solitonic objects as super-selection sectors of the perturbative string theory.
349: 
350: The moduli spaces of the string theory is connected by various dualities. In particular, one of the most important recent achievements is the advent of the gauge - gravity correspondence. Before this new development, it had been believed that the local quantum field theory cannot realize the gravitational theory (Weinberg-Witten theorem \cite{Weinberg:1980kq}). However, the holographic realization of the gauge theory avoid this no-go theorem in a remarkable manner, and it has enabled us to study the strongly coupled gauge theory from the weakly coupled gravity. Explicit realization in the string theory involves the low-energy decoupling limit (Maldacena limit \cite{Maldacena:1997re}) of the localized excitations: the most famous example is the low-energy field theory limit of open-string theory living on the D3-brane in flat ten-dimensional space, which yields the duality between type IIB string theory on $AdS_5 \times \mathbb{S}^5$ and the $\mathcal{N}=4$ supersymmetric Yang-Mills theory on $\br^{1,3}$ (or $\br^1 \times \mathbb{S}^3$) \cite{Maldacena:1997re,Aharony:1999ti}.
351: 
352: The decoupling limit of the localized degrees of freedom and the gauge - gravity correspondence are not only important for the understanding of the strongly coupled gauge theories, but also essential to understand the quantum gravitational nature of the string theory. What is the microscopic origin of the black hole entropy? What is the fundamental degrees of freedom for the quantum gravity? How does (or does not) string theory solve the information paradox? These questions have been partially answered from the gauge - gravity correspondence of D-branes.
353: The decoupling limit is essentially the near horizon limit of the corresponding supergravity background, and the properties of black hole can be understood through the gauge - gravity correspondence in this way.
354: 
355: For NS5-brane, situations are more involved. Compared with D-branes, the NS5-brane is more geometrical in its origin. Indeed, as we will see in section \ref{sec:2-2}, it is T-dual to the singular geometry, and it appears not obvious what is the localized degrees of freedom in the decoupling limit. On the other hand, the closed string background for the near horizon limit of the NS5-brane is exactly quantized, so we are able to understand the gauge - gravity correspondence beyond the supergravity approximation.
356: 
357: Our starting point is the supergravity solution for the extremal NS5-brane: the solution contains nontrivial dilaton and the metric\footnote{Throughout this thesis, we use the string frame for supergravity solutions.}
358: \begin{align}
359:  \dd s^2   \equiv G_{\mu\nu} \dd x^\mu \dd x^\nu =- \dd t^2
360: + \left(1+\frac{k\al'}{r^2}\right) \left(\dd
361: r^2+r^2 \dd \Omega_3^2\right)+ \dd {\bf
362: y}^2_{\mathbb{R}^5}~,  ~~~
363: e^{2\Phi(r)} = g_s^2 \left(1+\frac{k\al'}{r^2}\right) \ ,
364: \label{exNS5}
365: \end{align}
366: along with $k$-units of NS-NS $H_3$-flux penetrating through
367: $\mathbb{S}^3$:
368: \begin{align}
369: H_{mnp} = -\epsilon_{mnp}^{\ \ \ \ q} \partial_q \Phi(r) \ ,
370: \end{align}
371: where $x^m$ ($m=6,\cdots,9$) are transverse to the 5-brane.
372: Thus, $k$ refers to the number of NS5-branes at
373: $r=0$, ${\bf y}$ are
374: the spatial coordinates of the planar NS5-brane worldvolume, and
375: $g_s$ is the string coupling constant at infinity. The background preserves 16 supercharges of the type II (A or B) supergravity.
376: 
377: Following the argument of decoupling limit given above, we take the near horizon limit of the geometry \eqref{exNS5} by zooming in the $r^2 \ll \alpha'$ region. Neglecting the constant term (i.e. $1$) in the harmonic function $\left(1+\frac{k\al'}{r^2}\right)$, we obtain the near horizon limit of the extremal NS5-brane background \cite{Rey:1989xi,Rey:1989xj,Rey:1991uu,Callan:1991dj,Callan:1991at}
378: 
379: \begin{align}
380:  \dd s^2 = - \dd t^2 + k\al' \dd \rho^2 + k\al' \dd
381: \Omega_3^2 + \dd {\bf y}_{\mathbb{R}^5}^2 ~, \qquad {\Phi} = -\rho +
382: \text{constant}\ , \label{NH ext NS5}
383: \end{align}
384: where $r  = \sqrt{k\alpha'} \exp \rho$. This near horizon background remarkably admits an exact conformal field theory description involving a linear
385: dilaton theory and $SU(2)_k$ super Wess-Zumino-Novikov-Witten (WZNW)
386: model:\footnote{Here, $k$ is the level of total current of super SU(2) WZNW models and $\sqrt{\frac{2}{k}}$ is the amount of background charge for the linear dilaton system.}
387: %
388: \begin{align} \Big[ \mathbb{R}_t \times \mathbb{R}_{\rho , \sqrt{2\over k}}
389: \times SU(2)_k \Big]_\perp \times \Big[\mathbb{R}^5 \Big]_{||}\ .
390: \end{align}
391: %
392: The first part describes the five-dimensional curved space-time
393: transverse to the NS5-brane while the second part describes the
394: flat spatial directions parallel to the NS5-brane. The criticality
395: condition for superstring theories is satisfied for any $k$ because
396: %
397: \begin{align}
398:  \left( 1 + \frac{6}{k} + \frac{1}{2}\right) + 3 \times \left(
399: \frac{k-2}{k} + \frac{1}{2}\right)+6\times \left(1 + \frac{1}{2}
400: \right) =15 \label{crit} \ .
401: \end{align}
402: 
403: Although the background is exactly solvable, the string background is 
404: singular due to the existence of the linear dilaton direction $\rho$.
405: In the large negative $\rho$, the string coupling constant effectively diverges 
406: and the string perturbation theory is ill-defined. Physically, there exists a core of NS5-branes at $r=0$, and the dynamical degrees of freedom on the NS5-brane cannot be neglected.
407: 
408: There are several ways to regularize this linear dilaton singularity so that the string world-sheet perturbation theory makes sense with sufficient predictive power. 
409: One way to do this is to introduce the non-extremality to the geometry. Let us consider the non-extremal or black NS5-brane solution in the ten-dimensional type II supergravity:
410: \begin{align}
411:  \dd s^2 = -\left(1-\frac{r_0^2}{r^2}\right) \dd t^2
412: + \left(1+\frac{k\al'}{r^2}\right) \left(\frac{\dd
413: r^2}{1-\frac{r_0^2}{r^2}} +r^2 \dd \Omega_3^2\right)+ \dd {\bf
414: y}^2_{\mathbb{R}^5}~,  ~~~
415: e^{2\Phi(r)} = g_s^2 \left(1+\frac{k\al'}{r^2}\right) \ 
416: \label{blackNS5}
417: \end{align}
418: along with $k$-units of NS-NS $H_3$-flux penetrating through
419: $\mathbb{S}^3$ again. Here $r=r_0$ is the location of the event horizon of the black NS5-brane.
420: 
421: One type of near-horizon limit
422: is $r_0 \rightarrow 0$ and $g_s \rightarrow 0$ independently,
423: leading to the `throat geometry' of extremal NS5-branes that reduce to
424: \eqref{NH ext NS5}.
425: Another type of near-horizon limit is $r_0 \rightarrow 0$ and $g_s
426: \rightarrow 0$ while keeping the energy density above the extremal
427: configuration $\mu \equiv {r_0^2}/{g_s^2 \al'}$ fixed. It yields
428: `throat geometry' of the near-extremal NS5-branes \eqref{blackNS5}
429: \cite{Maldacena:1997cg,Kutasov:2000jp}:
430: \begin{align}
431: \hspace{-5mm} \dd s^2 = - \tanh^2\rho \, \dd t^2 + k\al' \dd
432: \rho^2 + k\al' \dd \Omega_3^2 + \dd {\bf y}_{\mathbb{R}^5}^2 ~,
433: \qquad e^{2\Phi} = \frac{k}{\mu \cosh^2 \rho} ~, \label{NH black
434: NS5}
435: \end{align}
436: %
437: where $r= r_0\cosh \rho$. For $(t,\rho)$-subspace, the metric and
438: the dilaton coincide with those of the two-dimensional black hole with a Lorentzian signature.
439: This Lorentzian black hole can be described by Kazama-Suzuki
440: supercoset conformal field theory $SL(2; \br)_k / U(1)$ (where
441: $U(1)$ subgroup is chosen to be the non-compact component
442: (i.e. space-like direction)) of central charge $c=3(1+2/k)$. Likewise,
443: taking account of the NS-NS $H_3$-flux penetrating through
444: $\mathbb{S}^3$ which is omitted in (\ref{NH black NS5}), the angular
445: part $\mathbb{S}^3$ can be described by the (super) $SU(2)$-WZNW model as we have seen in the extremal case.
446: In this way, the string background of the nonextremal NS5-brane is
447: reduced to a solvable superconformal field theory system:\footnote
448:   {Here again, $k$ denotes the level of total current of super WZNW models.
449:    Namely, $k+2$, $k-2$ are the levels of bosonic
450:     $SL(2)$ and $SU(2)$ currents.}
451: \begin{align}
452: \Big[ {SL(2;\br)_{k} \over U(1)} \times SU(2)_{k} \Big]_\perp \times
453: \Big[\, \mathbb{R}^5 \, \Big]_{||}~. \label{SCFT black NS5}
454: \end{align}
455: Here, the first part describes the five-dimensional curved space-time
456: (including the time direction) transverse to the NS5-brane, while
457: the second part describes the flat spatial directions parallel to
458: the NS5-brane. The criticality condition is satisfied for any $k$ as
459: in \eqref{crit}.
460: 
461: As we will review in the next section, the classical geometry of the two-dimensional black hole itself is not singularity free. This is because although in the Schwarzshild-like coordinate used in \eqref{NH black NS5} there is no singularity at all, the event horizon exists at $\rho=0$, and we can extend the coordinate inside the horizon. In the maximally extended geometry, we observe a curvature and dilaton singularity as is the case with the usual Schwarzshild black hole. It is interesting, however, despite the appearance of the singularity, the exact SCFT formulation \eqref{SCFT black NS5} appears perfectly well-defined, at least formally.
462: 
463: Another way to regularize the linear dilaton singularity, while keeping the space-time supersymmetry in contrast with the above non-extremal resolution, is to separate the position of the NS5-branes in a ring-like manner and study the smeared solution \cite{Sfetsos:1998xd} (see also \cite{Israel:2004ir,Itzhaki:2005zr}). The background is described by the coset model 
464: \begin{align}
465: \frac{\Big[ {SL(2;\br)_{k} \over U(1)} \times \frac{SU(2)_{k}}{U(1)} \Big]_\perp}{\bz_k} \times
466: \Big[\, \mathbb{R}^{1,5} \, \Big]_{||}~. \label{SCFT sNS5}
467: \end{align}
468: Here $\bz_k$ orbifold serves as a GSO projection\footnote{With the abuse of convention, the Gliozzi-Scherk-Olive (GSO) projection has a two-fold meaning in this thesis (and in many literatures). The one is the summation over the spin structure \cite{Gliozzi:1976qd}, and the other is the restriction to the integral $U(1)_R$ charge sector for the internal SCFT. Both are imperative to preserve the target-space supersymmetry.} that restricts the spectrum to the sector with integral $U(1)_R$ charge so that the space-time supercharge is well-defined. Intuitively, we have extracted a particular $U(1)$ direction from the $SU(2)$ WZNW model and combined it with the linear dilaton direction to construct  the Euclidean ${SL(2;\br)_{k} \over U(1)}$ coset model by a marginal deformation.\footnote{$U(1)$ subgroup here is chosen to be the compact direction.} The linear dilaton direction together with the $U(1)$ direction is deformed to the ${SL(2;\br)_{k} \over U(1)}$ coset model that does not possess a dilaton singularity.
469: 
470: To see the geometrical meaning of this deformation, we write the coset part $\Big[ {SL(2;\br)_{k} \over U(1)} \times \frac{SU(2)_{k}}{U(1)} \Big]_\perp $ as
471: \begin{align}
472: \dd s^2 = \alpha'k(\dd \theta^2 + \tan^2\theta \dd\tilde{\phi}^2_2 + \dd\rho^2 + \tanh^2\rho \dd\tilde{\phi}_1^2) \ ,  \ \ e^{2\Phi} = \frac{1}{\cos^2\theta\cosh^2\rho} \ . \label{startm}
473: \end{align}
474: It is interesting to note that this geometry does {\it not} admit any Killing spinor needed for an apparent supersymmetry: the supersymmetry will be recovered after taking the $\bz_k$ orbifold \cite{Israel:2004ir} (see also \cite{Bakas:1994ba,Bergshoeff:1994cb,Bakas:1995hc} for earlier discussions). The $\bz_k$ orbifold is defined as $(\tilde{\phi}_1,\tilde{\phi}_2) \sim (\tilde{\phi}_1+2\pi/k,\tilde{\phi}_2+2\pi/k)$. We define new coordinates 
475: \begin{align}
476: \tilde{\phi}_1 = \phi_1+\phi_2/k \ , \ \ \tilde{\phi}_2 = \phi_2/k
477: \end{align}
478: so that the $\bz_k$ orbifold simply acts as $(\phi_1,\phi_2) \sim (\phi_1, \phi_2 + 2\pi)$. In the new coordinates, the metric reads
479: \begin{align}
480: \dd s^2 = \alpha'k(\dd\theta^2 +\dd\rho^2 \tanh^2\rho \dd\phi_1^2) + 2\alpha'\tanh^2\rho \dd\phi_1\dd\phi_2 + \frac{\alpha'}{k}(\tan^2\theta + \tanh^2\rho) \dd\phi_2^2 \ .
481: \end{align}
482: Since $\phi_2$ direction has a usual $2\pi$ periodicity, one can perform the T-duality along the $\phi_2$ direction. Applying Buscher's rule (see appendix \ref{busc}), we obtain
483: \begin{align}
484: \dd s^2 &= \alpha'k\left(\dd\theta^2 + \dd\rho^2 + \frac{\tan^2\theta\tanh^2\rho}{\tan^2\theta+\tanh^2\rho} \dd\phi_1^2 + \frac{1}{\tan^2\theta + \tanh^2\rho}{\dd\hat{\phi}_2} \right)  \cr
485: B&= \frac{\alpha'k\tanh^2\rho}{\tan^2\theta+ \tanh^2\rho}\dd\phi_1\wedge \dd\hat{\phi}_2 \ , \ \ e^{2\Phi} = \frac{1}{\cos^2\theta\cosh^2\rho(\tan^2\theta+\tanh^2\rho)} \ . \label{rmet}
486: \end{align}
487: 
488: In the asymptotic region $\rho \to \infty$, we recover the NS5-brane solution \eqref{NH ext NS5}, and we can also see that the NS5-branes are now localized along the ring $\theta = \rho=0$, where the dilaton diverges (see figure \ref{fig:ring} for a description of our coordinate system). In other words, the NS5-branes are located along the ring in the $(x^8,x^9)$ plane.\footnote{Our parametrization is $x^6 = \rho_0 \sinh\rho \sin \theta \cos \phi_1$, $x^7 = \rho_0 \sinh\rho \sin \theta \sin\phi_1$, $x^8 = \rho_0 \cosh\rho\cos\theta \cos \hat{\phi}_2$, $x^9 = \rho_0\cosh\rho \cos \theta \sin\hat{\phi}_2$. } In this sense, the geometry still appears singular, but as we will discuss later, this is just an artefact of loose applications of T-duality: the trumpet singularity in \eqref{rmet} will be resolved by the ``winding tachyon condensation". Another quick way to see the absence of singularity is to revisit our starting point \eqref{startm}: it does not possess any dilaton singularity. It is also clear that the coset \eqref{SCFT sNS5} is manifestly singularity free as an SCFT up to a harmless orbifold structure.
489: 
490: Although we will not explicitly do it here, we can begin with the appropriate (smeared) harmonic function ansatz for the ring-likely distributed NS5-branes and reproduce the metric \eqref{rmet} purely from the supergravity solution by taking a suitable near horizon limit \cite{Sfetsos:1998xd}. In this approach, the space-time supersymmetry is manifest.
491: 
492: \begin{figure}[htbp]
493:    \begin{center}
494:     \includegraphics[width=0.5\linewidth,keepaspectratio,clip]{ring.eps}
495:     \end{center}
496:     \caption{NS5-branes are localized along the ring in the $(x_8,x_9)$ plane with $\theta = \rho = 0$.}
497:     \label{fig:ring}
498: \end{figure}
499: \subsection{Noncritical superstring and LST}\label{sec:2-2}
500: In section \ref{sec:2-1}, we discussed the relation between the two-dimensional black hole systems and the near horizon NS5-brane configurations. It is possible to generalize this construction to describe the singular limit of the geometry from exactly solvable conformal field theories. The construction is similar to the Gepner construction  for compact Calabi-Yau spaces \cite{Gepner:1987vz,Gepner:1987qi}, and it can be named ``non-compact Gepner construction" \cite{Ghoshal:1995wm,Ooguri:1995wj,Giveon:1999zm,Lerche:2000uy,Hori:2002cd,Eguchi:2004ik}. In this subsection, we would like to review this construction. Thanks to this generalized ``non-compact Gepner construction", most of the results we will present in later sections can be applied to various singular Calabi-Yau spaces.
501: 
502: \subsubsection{noncompact Calabi-Yau and wrapped NS5-branes}\label{sec:2-2-1}
503: Our starting points to construct exactly solvable world-sheet conformal field theories for singular Calabi-Yau spaces from two-dimensional black hole and minimal models are the following two claims.
504: 
505: {\bf Calabi-Yau / Landau-Ginzburg correspondence} \cite{Vafa:1988uu,Lerche:1989uy,Martinec:1988zu,Witten:1993yc}
506: 
507: Let us consider the algebraic varieties defined by
508: \begin{align}
509: \sum_{i=1}^{n+2} x_i^{r_i} = 0 \label{nocv}
510: \end{align}
511: in the weighted projective space $\mathbb{WCP}_{n+1}\left(\frac{1}{r_1},\cdots
512: ,\frac{1}{r_{n+2}}\right)$. The Calabi-Yau condition reads $\sum_{i=1}^{n+2}\frac{1}{r_i} = 1$. The Calabi-Yau / Landau-Ginzburg correspondence says that the (quantum) sigma model defined on the $n$-dimensional algebraic varieties \eqref{nocv} is (weakly) equivalent\footnote{The precise meaning of the weak equivalence can be found e.g. in \cite{Hori:2000kt}.} to the $\mathcal{N} = 2$ supersymmetric Landau-Ginzburg orbifold theory with the superpotential
513: \begin{align}
514: W(X_i) = \sum_{i=1}^{n+2} X_i^{r_i}  \ . \label{lgo}
515: \end{align}
516: The orbifold projection serves as a GSO projection demanding the integrality of the $U(1)_R$-charge of the total model. The Calabi-Yau condition can be understood as the criticality condition for the SCFT with the central charge $\hat{c} = c/3 = n$.
517: 
518: When all $r_i$ are positive, the resulting model is nothing but the Gepner construction for compact Calabi-Yau spaces (see also \cite{Greene:1988ut,Witten:1993yc}). When some of $r_i$ are negative, the Calabi-Yau manifold is non-compact and the definition of the Landau-Ginzburg orbifold needs extra care as we will do it momentarily.
519: 
520: {\bf Landau-Ginzburg / minimal model correspondence} \cite{Lerche:1989uy}
521: 
522: The $\mathcal{N}=2$ supersymmetric Landau-Ginzburg model with the superpotential $W(X) = X^{k}$ is equivalent to the $\mathcal{N}=2$ $(k-2)$-th minimal model with the central charge $\hat{c} = c/3 = 1 - \frac{2}{k}$. The minimal model has an algebraic formulation, but an alternative construction is based on the Kazama-Suzuki coset $SU(2)_k/U(1)$ associated with the level $k$ $SU(2)$ current algebra.\footnote{We always stick to the convention where $k$ denotes the {\it total} level of the current algebra: the bosonic $SU(2)$ current algebra has the level $\kappa = k-2$ and the bosonic $SL(2;\br)$ current algebra has the level $\kappa = k+2$.} Kazama-Suzuki construction guarantees that the coset CFT associated with the $\mathcal{N}=1$ current algebra actually possesses $\mathcal{N}=2$ superconformal symmetry when the target space is a special Kahler manifold (the hermitian symmetric manifold) \cite{Kazama:1988uz,Kazama:1988qp}. In our simplest case, it is indeed the case and the theory is equivalent to the $\mathcal{N}=2$ minimal model.\footnote{Actually, if the denominator $H$ in the coset $G/H$ is a Cartan subgroup of $G$, the coset admits the $\mathcal{N}=2$ superconformal symmetry even if it is a non-symmetric space \cite{Kazama:1988va}.}
523: 
524: We can formally generalize the above discussion to define the $\mathcal{N}=2$ supersymmetric Landau-Ginzburg model with the negative power superpotential $W(X) = X^{-k}$. The analytic continuation of the central charge for the positive power superpotential yields $\hat{c} = c/3 = 1 + \frac{2}{k}$. The Kazama-Suzuki coset construction has a natural generalization in this case as well. Instead of considering $SU(2)_k/U(1)$ supercoset model, we consider $SL(2;\br)_k/U(1)$ supercoset model whose central charge is also given by $\hat{c} = c/3 = 1 + \frac{2}{k}$. This CFT will be reviewed in section \ref{sec:3}. Since the Lagrangian formulation based on the Landau-Ginzburg model with the negative power superpotential does not seem to be well-defined while the $SL(2;\br)_k/U(1)$ coset does, the precise claim of the non-compact Gepner construction is that the Landau-Ginzburg orbifold appearing in \eqref{lgo} should be understood as the $SL(2;\br)_k/U(1)$ coset model.
525: 
526: At this point, it would be interesting to mention that the formal Landau-Ginzburg description suggests a duality between the $\mathcal{N}=2$ Liouville theory and the $SL(2;\br)_k/U(1)$ coset model. We begin with the topological path integral for the partition function on the sphere: 
527: \begin{align}
528: Z &= \int \dd X \dd \bar{X} \frac{1}{g_s^2} e^{-W(X)-\bar{W}(\bar{X})}  \cr
529: &=\int \dd X \dd \bar{X} \frac{1}{g_s^2}e^{-X^{-k}-\bar{X}^{-k}} \ . \label{pi}
530: \end{align}
531: Introducing the $\mathcal{N}=2$ Liouville coordinate $X^{-k} = e^{\frac{1}{\mathcal{Q}}\Phi}$ with $\mathcal{Q}^2 = \frac{2}{k}$, we can rewrite the path integral \eqref{pi} as
532: \begin{align}
533: Z = \int \dd\Phi \dd\bar{\Phi} \frac{1}{g_s^2} \exp\left({-\mathcal{Q}\mathrm{Re}\Phi -e^{\frac{1}{\mathcal{Q}}\Phi}- e^{\frac{1}{\mathcal{Q}}\bar{\Phi}}}\right) \ .
534: \end{align}
535: The important step is to regard the measure factor $\exp\left({-\mathcal{Q}\mathrm{Re}\Phi}\right)$ as a space-dependent coupling constant, namely, linear dilaton background with the slope $\mathcal{Q}$. The remaining action shows the structure of the $\mathcal{N}=2$ Liouville superpotential $W(\Phi) = e^{\frac{1}{\mathcal{Q}}\Phi}$. This heuristic equivalence between the $SL(2;\br)/U(1)$ coset model and the $\mathcal{N}=2$ Liouville theory at the topological level will be made more precise in later section \ref{sec:3-4}.
536: 
537: Now combining these two facts, we can construct the equivalent description for  (non-compact) Calabi-Yau spaces by considering tensor products of $SL(2;\br)/U(1)$ coset models ($\mathcal{N}=2$ Liouville theory) and $SU(2)/U(1)$ coset models ($\mathcal{N}=2$ minimal models) with appropriate GSO projections. We call such a construction a generalized (non-compact) Gepner construction.
538: 
539: Let us discuss some simple examples.
540: 
541: 1) $A_{k-1}$ type ALE spaces
542: 
543: We take $n=2$, and set $-r_1 = r_2 = k$, $r_3=r_4=2$. From the projective invariance, we can set $x_1= - \mu$ without loss of generality.\footnote{Note that we are considering a noncompact space, so the domain of the projective coordinate $x_1$ is in $\mathbb{C}^* \equiv \mathbb{C} - \{0\}$. } The resulting algebraic variety is given by
544: \begin{align}
545: x_2^k + x_3^2 + x_4^2 = \mu^k \label{deak}
546: \end{align}
547: in $\mathbb{C}^3$, which is nothing but the $A_{k-1}$ type ALE space with a deformation parametrized by $\mu$. On the other hand, the noncompact Gepner construction yields
548: 
549: \begin{align}
550: \frac{\Big[ {SL(2;\br)_{k} \over U(1)} \times \frac{SU(2)_{k}}{U(1)} \Big]}{\bz_k}
551: \end{align}
552: because the massive theory with the quadratic superpotential $W(X) = X^2$  will decouple under the renormalization group flow. We now recognize that the resulting theory is same as the near horizon limit of the $k$ NS5-brane solutions discussed in section \ref{sec:2-1}. This shows an equivalence between the near horizon limit of $k$ NS5-brane solutions and the $A_k$ type ALE spaces. They are indeed related with each other via the T-duality. The deformation parameter $\mu$ in the ALE space corresponds to the separation of NS5-branes. We can easily generalize the construction for other ALE spaces with ADE singularities.
553: 
554: 2) Generalized conifolds
555: 
556: We next consider the case of Calabi-Yau three-fold $(n=3)$. We take $r_2,r_3,r_4,r_5> 0$ and set $r_1 = 1-\frac{1}{\sum_{i=2}^5 r_i^{-1}} < 0$. After fixing the projective invariance by eliminating $x_1$, the resultant Calabi-Yau space is the so-called generalized (deformed) conifold  
557: \begin{align}
558: x_2^{r_2} + x_3^{r_3} + x_4^{r_4} + x_5^{r_5} = \mu \ \label{dgc}
559: \end{align}
560: in $\bc^4$. Mathematically, we can regard it as a complex structure deformation of the Brieskorn-Pham type singularity (see section \ref{sec:2-2-3}). The Gepner construction leads to the orbifolded tensor products of $\mathcal{N}=2$ minimal models with one $SL(2;\br)_{-r_1}/U(1)$ coset model.
561:  The simplest example is the case with $r_1=-1$ and $r_2 = r_3= r_4=r_5=2$. The geometry is the deformed conifold:
562: \begin{align}
563: x_2^2 + x_3^2 + x_4^2 + x_5^2 = \mu \ .
564: \end{align}
565: The noncompact Gepner construction is given by $SL(2;\br)_1/U(1)$ coset model with the level 1 parent current algebra. This is the famous Ghoshal-Vafa duality \cite{Ghoshal:1995wm}.
566: 
567: 3) ALE($A_{k-1})$ fibration over weighted projective spaces
568: 
569: We finally consider the model with two negative charges: $n=3$, $r_1=-k(1+\frac{k_1}{k_2})$, $r_2 = -k(1+\frac{k_2}{k_1})$, $r_3 = k$, and $r_4=r_5=2$. The Landau-Ginzburg superpotential is given by
570: \begin{align}
571: W(X_i) = X_1^{-k(1+\frac{k_1}{k_2})} + X_2^{-k(1+\frac{k_2}{k_1})} + X_3^{k} \ .
572: \end{align}
573: By introducing new variables: $Z= \log X_1 + \log X_2$, $Y=kk_1\log X_1 - kk_2 \log X_2$ and $X^k = e^{kZ} X_3^k $, we can rewrite the Landau-Ginzburg superpotential as
574: \begin{align}
575: W = e^{-kZ}(X^k + e^{Y/k_1} + e^{-Y/k_2}) \ ,
576: \end{align}
577: with the linear dilaton $\Phi = -\mathrm{Re} Z$. After integrating over $Z$, the topological path integral is localized along the locus\footnote{We have added the superpotential term $W_1^2 + W_2^2$ by hand to match the dimensionality.} 
578: \begin{align}
579: e^{y/k_1} + e^{-y/k_2} + x^k + w_1^2 + w_2^2 = 0 \ ,
580: \end{align}
581: which shows a structure of ALE($A_{k-1}$) fibration over 
582:  $\mathbb{WCP}^1(k_1,k_2)$. The simplest example with $k_1=k_2$, we obtain the ALE($A_{k-1}$) fibration over $\mathbb{CP}_1$. The geometry of the two $SL(2;\br)/U(1)$ coset and one $SU(2)/U(1)$ coset can be analysed in a similar way as we did in section \ref{sec:2-1}, and the result is given by the wrapped NS5-brane solution around $\mathbb{CP}_1$, where we have chosen $k_1=k_2=1$ for simplicity (see \cite{Hori:2002cd} for details). This is expected from the fact that the $A_{k-1}$ singularity is T-dual to flat $k$ NS5-branes and we could perform the fiber-wise T-duality for the ALE($A_{k-1}$) fibration over $\mathbb{CP}_1$.
583: 
584: The partition functions and elliptic genera of these noncompact Gepner models have been studied in \cite{Eguchi:2004yi,Eguchi:2004ik,Eguchi:2006tu}. In this section we restricted ourselves to the Landau-Ginzburg construction where the theory is defined as (an orbifold of) the direct product of the Landau-Ginzburg models with monomial superpotentials. Geometrically, they corresponded to the (deformations of) the Brieskorn-Pham type singularities. It is possible to construct more general Landau-Ginzburg orbifolds with generic polynomial superpotentials. The generalized models have a potential applications to the singular locus of the $\mathcal{N}=2$ supersymmetric Yang-Mills theories (Argyres-Douglas point) and their deformations. The exact quantization of the world-sheet theory beyond the topological subsector, however, is a difficult task and we would not pursue these generalizations any further in this thesis.
585: 
586: 
587: \subsubsection{singular limit and LST}\label{sec:2-2-2}
588: In section 2.1, we have discussed that the coinciding $k$ NS5-branes superstring solution corresponds to the linear dilaton background while the supersymmetric deformation (separation of NS5-branes in a ring-like manner) corresponds to the $SL(2;\br)/U(1)$ coset background (i.e. two-dimensional Euclidean black hole). Here we would like to take the similar singular limit in more general noncritical superstring backgrounds discussed in section \ref{sec:2-2-1}.\footnote{What we mean by ``noncritical" here is that the SCFTs involved does not necessarily possess an apparent 10-dimensional background as is the case with the Gepner construction for compact Calabi-Yau spaces. In a more specific narrower sense, we sometimes call a theory ``noncritical" when it possesses a Liouville direction.}
589: 
590: It is particularly easy to see the singular limit if we start with the $\mathcal{N}=2$ Liouville description: it has a superpotential
591: \begin{align}
592: W(\Phi) = \mu e^{\frac{1}{\mathcal{Q}}\Phi} \ ,
593: \end{align}
594: and the parameter $\mu$ directly corresponds to the deformation parameter appearing e.g. in \eqref{deak},\eqref{dgc}. Thus the singular limit $\mu \to 0$ is equivalent to switching off the Liouville potential so that we are left with the linear dilaton theory. The duality between the $\mathcal{N}=2$ Liouville theory and $SL(2;\br)/U(1)$ coset theory then confirms the statement that the singular limits of the non-compact Gepner models correspond to replacing $SL(2;\br)/U(1)$ coset part by the $\mathcal{N}=2$ supersymmetric linear dilaton theory with the same central charge and the same asymptotic dilaton gradient.
595: 
596: Let us formulate the proposal discussed above in a more precise way \cite{Giveon:1999zm}. We begin with the type II string theory on a singular Calabi-Yau varieties $X^{2n}$ with  the complex dimension $n$ defined as the vicinity of a hypersurface singularity 
597: \begin{align}
598: F(z_1,\cdots, z_{n+1}) = 0 \label{defeq}
599: \end{align}
600: in $\mathbb{C}^{n+1}$, where $F$ is a quasi-homogeneous polynomial on $\mathbb{C}^{n+1}$. This means that $F$ has degree $1$ under the $\mathbb{C}^{*}$ action:
601: \begin{align}
602: z_i \to \lambda^{r_i} z_i \ . \label{Cac}
603: \end{align}
604: 
605: Now we can define a locally holomorphic $n$-form $\Omega$ as
606: \begin{align}
607: \Omega = \frac{\dd z_1\wedge \dots \wedge \dd z_n}{\partial F/\partial z_{n+1}} \ 
608: \end{align}
609: on the patch $\partial F/\partial z_{n+1}\neq 0$. It can be extended to other patches where $\partial F\partial z_i \neq 0$ with the similar expressions and glued together to form a globally well-defined holomorphic $n$-form with the charge $r_{\Omega} = \sum_i r_i -1$ under the $\mathbb{C}^{*}$ action \eqref{Cac}. Such constructed varieties $X^{2n}$ are Gorenstein\footnote{Gorenstein means that  $X^{2n}-\{0\}$ admit a nowhere vanishing holomorphic $n$-form.} equipped with a natural $\mathbb{C}^{*}$ action \eqref{Cac} by construction.
610: 
611: We consider the type II string theory on the singular Calabi-Yau varieties $\br^{d-1,1}\times X^{2n}$ in the vicinity of the isolated singular point $y_0$ in the decoupling limit $g_s \to 0$. The proposed dual theory is the type II string theory on $\br^{d-1,1} \times \br_{\phi} \times \mathcal{N}$. Here $\mathcal{N}$ is the infrared limit of the sigma model on the manifold $\mathcal{N} = X^{2n}/\br_+$, where the division by $\br_+$ is an action on $z_i$ as \eqref{Cac} with $\lambda \in \br_+$, The infrared limit\footnote{We assume $r_{\Omega}>0$ so that $\mathcal{N}$ is Fano meaning that the curvature of the Einstein metric on it is positive.} of the sigma model on $\mathcal{N}$ is given by a Landau-Ginzburg model with superpotential $F(Z_i)$ and an additional $\mathbb{S}^1$ circle.\footnote{As a CFT, they are not a simple direct product but an orbifold. We need to impose the GSO projection to preserve the target-space supersymmetry.} Here $\mathbb{S}^1$ direction corresponds to the $U(1)_R$ symmetry associated with the residual $U(1)$ action \eqref{Cac} with $|\lambda| =1$. The quotient space $\mathcal{N}/U(1)$ is equivalent to the Landau-Ginzburg model with superpotential $W= F(Z_i)$ from the standard Landau-Ginzburg / non-linear sigma model correspondence. The linear dilaton slope is determined by the total criticality condition of the string theory.
612: 
613: 
614: This construction is equivalent to the one discussed above by turning off the $\mathcal{N}=2$ Liouville superpotential or $SL(2;\br)/U(1)$ deformation. For later purposes, it is worthwhile studying the normalizability of such deformations.
615:  Consider the variation of the complex structure of $X^{2n}$:
616: \begin{align}
617: F(z_i) + \sum_a t_a A_a(z_i)  = 0 \ .
618: \end{align}
619: Here $t_a$ are complex deformation parameters, and $A_a(z_i)$ are complex structure deformations of the defining equation \eqref{defeq}. The Kahler potential of the Weil-Petersson metric for such complex structure deformations is given by the formula \cite{Candelas:1990pi}
620: \begin{align}
621: K = - \log \int_{X^{2n}} \Omega \wedge \bar{\Omega} \ .
622: \end{align}
623: To discuss the normalizability of the perturbation associated with $A_a$, we have to evaluate
624: \begin{align}
625: \frac{\partial^2}{\partial t_a \partial \bar{t}_a} \Omega \wedge \bar{\Omega} \ . \label{metd}
626: \end{align}
627: This can be done by the simple scaling argument \cite{Gukov:1999ya}: if $A_a$ scales under \eqref{Cac} as $\lambda^{r_a}$, $t_a$ should scale as $\lambda^{1-r_a}$, so \eqref{metd} scales with a weight 
628: \begin{align}
629: \omega_a = 2\left(\sum_{i}r_i + r_a - 2\right) \ .
630: \end{align}
631: The modes satisfying $r_a > 1-r_{\Omega}$ are non-normalizable deformations as $|z_i| \to \infty$ while the modes satisfying $r_a < 1-r_{\Omega}$ are normalizable deformations. 
632: 
633: The non-normalizable deformations should be regarded as boundary conditions we have to impose at infinity to define a theory. Different boundary conditions would give rise to different theories. On the other hand, the normalizable deformations should be regarded as fluctuating fields after the quantization. Their values can be varied within a given theory. As we will discuss later in section \ref{sec:3-1}, the normalizability of the deformations from the space-time viewpoint presented here is deeply connected with the normalizability of the corresponding operators in the world-sheet $\mathcal{N}=2$ linear dilaton theory. Indeed, one can regard this agreement as a nontrivial support for the duality proposed in \cite{Giveon:1999zm} and reviewed here.
634: 
635: As an example, let us consider a class of generalized conifolds defined by the hypersurface
636: \begin{align}
637: F(z_i) = H(z_1,z_2) + z_3^2 + z_4^2 \ 
638: \end{align}
639: in $\bc^4$. It can be regarded as an NS5-brane wrapped around the Riemann surface $H(z_1,z_2)= 0$ along the line of arguments reviewed at the end of section \ref{sec:2-2-1}. We begin with the $A_{n-1}$ type Brieskorn-Pham singularity with $H(z_1,z_2) = z_1^n + z_2^2$, and consider the perturbations of the form $z_1^a$ $(a=0,1,\cdots,n-2)$. $U(1)_R$-charges are given by $r_{\Omega} = \frac{1}{n}+\frac{1}{2}$ and $r_a = \frac{a}{n}$. From the condition $r_a > 1-r_{\Omega}$, we conclude that the deformations with
640: \begin{align}
641: a > \frac{n}{2}-1 \label{normc}
642: \end{align}
643: are non-normalizable \cite{Gukov:1999ya,Giveon:1999zm}. We can also understand the normalizability of these deformations from the dual $\mathcal{N}=2$ supersymmetric four-dimensional field theory viewpoints by studying the Seiberg-Witten theory near the Argyres-Douglas points \cite{Argyres:1995jj,Argyres:1995xn,Eguchi:1996vu}. 
644: 
645: To conclude this section, we would like to revisit the question: what is the decoupling limit of the theory defined on the singularities \eqref{defeq}? We have reviewed the proposed {\it dual} string theory defined as (deformations of) the $\mathcal{N}=2$ linear dilaton theory coupled with the Landau-Ginzburg model. We here summarize the low-energy decoupled physics from the original NS5-brane construction in flat ten-dimensional Minkowski space. The decoupled theory has a conventional name ``little string theory (LST)"  \cite{Losev:1997hx}.\footnote{See \cite{Aharony:1999ks} for an earlier review.}
646: 
647: \begin{itemize}
648: 	\item As a decoupled six-dimensional theory, it has $\mathcal{N}=(2,0)$  (type IIA) or $\mathcal{N}=(1,1)$ (type IIB) supersymmetry. The theory is non-local.
649: 	\item They are classified by the ADE classification.
650: 	\item IR limit is the six-dimensional super Yang-Mills theory in type IIB and the six-dimensional interacting (2,0) SCFT in type IIA \cite{Seiberg:1997ax}.
651: 	\item BPS excitation includes a string with tension $l_s$ (little string) and the theory shows a Hagedorn-like thermodynamics with the Hagedorn temperature $\beta_{\mathrm Hg} \sim 2\pi\sqrt{2k}$ (see section \ref{sec:2-4}). Most of these high-energy states are nonperturbative in nature.
652: \end{itemize}
653: 
654: After compactification (by wrapping NS5-branes on projective spaces for instance), we have four- (or two-) dimensional effective theory, which includes Seiberg-Witten theory near the Argyres-Douglas singularities, corresponding to the Calabi-Yau 3-fold singularities.
655: 
656: The properties of these theories, known as the LST, are less known than the field theory living on D-branes. However, the closed string dual theory is  exactly quantizeable in many cases unlike the R-R background in the near horizon limit of the D-branes. The study of $\alpha'$ exact information is an interesting subject of its own, besides the application to the dual theories, and we will pursue this direction in the following sections.
657: 
658: \subsubsection{obstruction for conical metrics}\label{sec:2-2-3}
659: So far, we have assumed the existence of the Calabi-Yau varieties defined on the hypersurface singularity \eqref{defeq}. In the compact Calabi-Yau case such as the curve defined in \eqref{nocv}, Calabi-Yau theorem guarantees the existence of the unique Ricci-flat Kahler metric once the Calabi-Yau condition is satisfied. The existence of the Calabi-Yau metric on the hypersurface singularities \eqref{defeq}, however, is an open problem (see \cite{SE} for a review).
660: 
661: From the $\mathbb{C}^{*}$ action \eqref{Cac} on the complex variables $z_i$, it is natural to assume the conical metric
662: \begin{align}
663: \dd s^2_{X^{2n}} = \dd r^2  + r^2 \dd s^2_L \ ,\label{metc}
664: \end{align}
665: where $r\le 0$ denotes the radial direction $\br_+$ and $\dd s^2_L$ is the Sasaki-Einstein metric of the link $L$ associated with the non-compact Calabi-Yau variety $X^{2n}/\{0\} = \br_+ \times L $. The Sasaki condition is equivalent to the statement that the total metric is Kahler, and the Einstein condition is equivalent to the statement that the total metric is Ricci flat.
666: 
667: It turns out to be extremely difficult to give necessary and sufficient conditions for the existence of such conical metric (or alternatively existence of the Sasaki-Einstein metric on the link $L$) while infinitely many examples of explicit metrics have been constructed quite recently \cite{Gauntlett:2004yd,Cvetic:2005ft,Cvetic:2005vk}.
668: 
669: For definiteness we restrict ourselves to the Brieskorn-Pham type singularities defined by the particular polynomial
670: \begin{align}
671: F(z_i) = \sum_{i=1}^{n+1} z_i^{a_i} \ . \label{bpsin}
672: \end{align}
673: The corresponding hypersurface singularities $X^{2n}$ are always isolated and Gorenstein. However from the following physical reasoning, we believe that not every singularity possesses a conical metric.
674: 
675: Assuming the existence of such a conical metric, we can compute the volume of such a hypothetical Sasaki-Einstein link $L$ by the formula \cite{Bergman:2001qi}
676: \begin{align}
677: \mathrm{Vol}(L) = \frac{r_{\Omega}^n}{n^n\prod_{i=1}^{n+1}r_i} \mathrm{Vol}(\mathbb{S}^{2n-1}) \ ,
678: \end{align}
679: where $\mathrm{Vol}(\mathbb{S}^{2n-2}) = \frac{2\pi^n}{(n-1)!}$.\footnote{We have assumed that the Reeb-vector (conformal $U(1)_R$-symmetry) is given by the natural $\mathbb{C}^{*}$ action \eqref{Cac}. If this is not the case, we have to determine the ``correct" Reeb-vector by using the $Z$-minimization \cite{Martelli:2005tp,Martelli:2006yb} ($a$-maximization \cite{Intriligator:2003jj}) principle. }
680: Via the AdS-CFT correspondence, the central charge $a$ of the dual SCFT living on D3-branes placed at the tip of the cone is related to the volume  of the Sasaki-Einstein link $L$ \cite{Gubser:1998vd} as 
681: \begin{align}
682: a \propto \frac{1}{\mathrm{Vol}(L)} \ .
683: \end{align}
684: On the other hand, the conjectured $a$-theorem states that $a$ is a decreasing function during a  renormalization group flow from UV to IR. Geometrically speaking, the relevant deformations to \eqref{bpsin} should increase the volume.\footnote{From the pure gravity viewpoint, this is a consequence of the weaker energy condition \cite{Freedman:1999gp}.} However, this statement is clearly violated in some explicit examples such as the series $F = z_1^k + z_2^2 + z_3^2 + z_4^2$ with $k>4$, where $\mathrm{Vol}(k+1) > \mathrm{Vol}(k)$ contradicting with the $a$-theorem.
685: 
686: Recently, \cite{Gauntlett:2006vf} has given two mathematical obstructions for the existence of conical Calabi-Yau metric for such varieties.
687: 
688: {\bf The Bishop obstruction}
689: 
690: For the existence of the conical Calabi-Yau metric \eqref{metc}, $\mathrm{Vol}(L) < \mathrm{Vol}(\mathbb{S}^{2n-1})$ is necessary. From the dual gauge theory viewpoint,  the condition corresponds to the fact that by appropriate Higgsing that decreases $a$, we can reach $\mathcal{N}=4$ SYM theory.
691: 
692: {\bf The Lichnerowicz obstruction}
693: 
694: When $X^{2n}$ admits a holomorphic function with $U(1)_R$-charge $\lambda <1$, the conical Calabi-Yau metric does not exist. For the Brieskorn-Pham type singularities, it is equivalent to the statement $r_{\Omega} \le n r_a$ for any deformation. From the dual gauge theory viewpoint, the condition corresponds to the unitarity bound of dual operators for the deformations. It can be shown that the Lichnerowicz obstruction also eliminate a possible violation of $a$-theorem for the Brieskorn-Pham type singularities (see appendix \ref{Lic}).
695: 
696: As an example let us consider the Calabi-Yau four-fold defined by 
697: \begin{align}
698: F = z_1^k + z_2^2 + z_3^2 + z_4^2 + z_5^2 = 0 \ . \label{fourc}
699: \end{align}
700: The conical Calabi-Yau metric only exists for $k=2$, and other varieties are obstructed from the Lichnerowicz bound. Thus the $AdS_3\times L_7$ compactification of M-theory is only possible for $k=2$. This should be so because otherwise the $a$-theorem would be violated or the weaker energy-condition would be spoiled. 
701: 
702: On the other hand, one can consider the two-dimensional string compactification on such a hypothetical conical Calabi-Yau manifold \eqref{fourc} and add $N_f$ fundamental strings on the noncompact $\br^{1,1}$ directions at the tip of the cone. Due to the gravitational backreaction, we can see the near horizon geometry would be $AdS_3 \times \mathcal{N}$ with the constant string coupling $g_s^2 \sim \frac{1}{N_f}$, where $\mathcal{N}$ has been introduced in section \ref{sec:2-2-2} denoting the infrared limit of the sigma model on the hypothetical Sasaki-Einstein link $L$ associated with \eqref{fourc}. As discussed in this section, the Sasaki-Einstein link $L$ is obstructed, but the string theory on $AdS_3  \times \mathbb{S}^1 \times (LG(F)\sim \mathcal{N}/U(1))$ has a well-defined perturbative description based on the $SL(2;\br)$ current algebra with level $k$ times Landau-Ginzburg orbifold defined by $F(Z_i)$ (or $\mathcal{N}=2 $ minimal model) \cite{Aharony:1999ti}.\footnote{This does not mean the existence of such Sasaki-Einstein metric because we are sitting at the Gepner-point of the sigma model, where the geometrical description is questionable due to large $\alpha'$ corrections.} Interestingly, unlike the $a$-theorem associated with the M-theory compactification on $AdS_3 \times \mathcal{N}$, the $c$-theorem for the dual CFT of $AdS_3 \times \mathbb{S}^1 \times (LG(F)\sim \mathcal{N}/U(1))$ is always satisfied because the dual CFT central charge, which is determined from the curvature of the $AdS_3 \sim SL(2;\br)$, is given by $c=6kQ_1$.\footnote{For instance, $k=\frac{n}{n+1}$ for $A_{n-1}$ type Calabi-Yau four-folds.}
703: 
704: 
705:  In a similar fashion, the near horizon geometry of  every Brieskorn-Pham singularities admit the noncritical string construction based on the non-compact Gepner models as we have presented in this section irrespective of the above-mentioned obstructions. It would be interesting to understand the obstructions of the existence of conical metrics for such singularities from the noncritical string theory viewpoint. For instance, we can translate the Lichnerowicz obstruction as the claim that every relevant deformations up to $z_i^{a_i-3}$ should be normalizable.
706: 
707: 
708: 
709: \subsection{Classical two-dimensional black hole}\label{sec:2-3}
710: In section \ref{sec:2-1}, we have introduced the two-dimensional black hole geometry as a near horizon limit of the black NS5-brane solutions in the type II superstring theory:
711: \begin{align}
712: \dd s^2 = k\al'(- \tanh^2\rho \, \dd t^2 +  \dd 
713: \rho^2) \ , \label{tbhm}
714: \end{align}
715: with nontrivial dilaton gradient $e^{2\Phi} = \frac{k}{\mu \cosh^2 \rho}$.\footnote{We have rescaled the normalization of $t$ for simplicity of notation. We will sometimes do this in the following without notice, for it would be convenient to stick to $2\pi$ periodicity in the Euclidean time direction after the Wick rotation.}
716: It has been claimed in the literature that the background is $\alpha'$ exact perturbatively in the type II superstring theory while the bosonic two-dimensional black hole might receive perturbative world-sheet $\alpha'\sim \frac{1}{k}$ corrections \cite{Tseytlin:1992ri}. We will discuss physical importance of the nonperturbative corrections later in section \ref{sec:3-4}.
717: 
718: In this subsection, we review the classical geometry of the two-dimensional black hole. First of all, the metric \eqref{tbhm} has an event horizon at $\rho = 0$, but the $(t,\rho)$ coordinate does not cover the whole causal region of the black hole. One can maximally extend the geometry \eqref{tbhm} by introducing the Kruscal coordinate
719: \begin{align}
720: u = \sinh\rho e^{t} \ , \ \ v= -\sinh\rho e^{-t} \ , \ \ \dd s^2 = -2k\frac{\dd u\dd v}{1-uv} \ , \ \  e^{2\Phi} = \frac{k}{\mu(1-uv)} \ .
721: \end{align}
722: Note that in two-dimension, it is always possible to introduce the conformal coordinate $(u,v)$ locally with the conformally flat metric $\dd s^2 = f(u,v)\dd u\dd v$.
723: In this coordinate, the event horizon is located at $uv = 0$, and inside the horizon, we encounter singularity at $uv=1$, where the curvature and the dilaton diverges. The Kruscal diagram can be found in figure \ref{fig:kruscal}. Causal region of the
724: Lorentzian black hole background has four boundaries: past and
725: future horizons ${\cal H}^\pm$, and past and future asymptotic
726: infinities ${\cal I}^\pm$.
727: 
728: \begin{figure}[htbp]
729:    \begin{center}
730:     \includegraphics[width=0.5\linewidth,keepaspectratio,clip]{kruscal.eps}
731:     \end{center}
732:     \caption{Kruscal diagram for the two-dimensional black hole system.}
733:     \label{fig:kruscal}
734: \end{figure}
735: 
736: 
737: We can also study the global structure of the metric by using the Penrose coordinates, and one can write down the Penrose diagram (see figure \ref{fig:bh}) of the two-dimensional black hole system, which looks exactly same as that for the four-dimensional Schwarzshild black hole system (upon neglecting $\mathbb{S}^2$ angular directions). This is one of the motivations to study the two-dimensional black hole system as an exactly solvable toy model for four-dimensional Schwarzshild black hole.
738: 
739: 
740: \begin{figure}[htbp]
741:    \begin{center}
742:     \includegraphics[width=0.5\linewidth,keepaspectratio,clip]{bh.eps}
743:     \end{center}
744:     \caption{Penrose diagram for the two-dimensional black hole.}
745:     \label{fig:bh}
746: \end{figure}
747: 
748: 
749: The geodesic motion of a particle with the minimal coupling interaction to the geometry
750: \begin{align}
751: S = \int \dd s \  = \int \dd\tau \sqrt{\frac{\dot{u}\dot{v}}{1-uv}} \ ,
752: \end{align}
753: where dot denotes the derivative with respect to $\tau$, is given  by
754: \begin{align}
755: \begin{cases} \ddot{u}(1-uv) = -v\dot{u}^2 \cr
756: \ddot{v}(1-uv) = -u \dot{v}^2 \ .
757: \end{cases}
758: \end{align}
759: Later, we will compare this with the string motion and D-particle motion, both of which show quite distinct properties.
760: 
761: Finally, we would like to study the ``mass" of the two-dimensional black hole.
762:  For this purposes, it is convenient to use the Schwarzshild(-like) coordinate \eqref{blackNS5}:
763: \begin{align}
764: \dd s^2 = -\left(1-\frac{2M}{r}\right)\dd t^2 +  \frac{k}{1-\frac{2M}{r}}\frac{\dd r^2}{r^2} \ ,  \ \ e^{2\phi} = r \ ,
765: \end{align}
766: where $ r_0 = 2M$ is the location of the horizon. From the expression, it is clear that we can easily shift the value of $M$ multiplicatively $M\to aM$ by the scaling of $r$ as $r \to r/a$. Therefore the physical meaning of the ``mass" of the two-dimensional black hole is solely determined by the value of the string coupling constant (dilaton) at the horizon $r=2M$. It corresponds to the fact that the mass parameter $M$ is related to the world-sheet $\mathcal{N}=2$ Liouville cosmological constant $\mu$ in the dual $\mathcal{N}=2$ Liouville theory, where $\mu$ can be shifted by the shift of the Liouville coordinate (see section \ref{sec:3-4} for more about the duality).  
767: 
768: Because of this property, the Hawking temperature of the two-dimensional black hole is independent of $M$ unlike the case with higher-dimensional black-holes. Similarly many features of the string theory in the two-dimensional black hole background such as scattering amplitudes also show rather trivial dependence on $M$.\footnote{It is customary to set $M=1$ as we will do in most part of the thesis.} It is related to the Knizhnik-Polyakov-Zamolodchikov (KPZ) scaling law of the dual Liouville theory \cite{Knizhnik:1988ak}.
769:  
770: \subsection{Wick rotation: thermodynamic properties}\label{sec:2-4}
771: 
772: In this section, we would like to study thermodynamic properties of the two-dimensional black hole (and hence LST on the black NS5-branes). It is a well-known but profound fact that the black hole system shows a thermodynamic properties \cite{Bekenstein:1973ur,Bardeen:1973gs}:
773: \begin{itemize}
774: 	\item Surface gravity $\kappa$ (temperature $T$) is constant over horizon of stationary black hole (the zeroth law: $T=\frac{\kappa}{2\pi}$)
775: 	\item $\dd M = \frac{1}{8\pi} \kappa \dd A + \Omega \dd Q$ (the first law: $S= \frac{A}{4}$)	
776: 	\item $\delta A \ge 0$ in any physical process (the second law)
777: 	\item It is impossible to achieve $\kappa = 0$ by any physical process (the third law)
778: \end{itemize}
779: Here $M$ is the mass of the black hole, $A = 4S$ is the area of the event horizon ($=$ entropy), $Q$ is the charge, and $\Omega$ is its chemical potential.
780: One of the biggest motivations to study the quantum gravity such as the string theory is to understand the black hole thermodynamics from the microscopic viewpoint.
781: 
782: Let us begin with the temperature of the two-dimensional black hole. One convenient way to compute the Hawking temperature of black hole systems \cite{Hawking:1974sw}
783:  is to use the Euclidean path integral formalism \cite{Gibbons:1976ue}. In our case, we can study the Wick rotation (i.e. $ t\to i\tau_E$) of the Lorentzian two-dimensional black hole \label{tbhm}:
784: \begin{align}
785: \dd s^2_E =  \tanh^2\rho \, \dd \tau_E^2 + k\al' \dd 
786: \rho^2 \ . \label{tbhmE}
787: \end{align}
788: To avoid a conical singularity at the origin $\rho=0$, we have to set the periodicity of the Euclidean time direction by $\beta_{\rm Hw} = {2\pi}{\sqrt{k\alpha'}}$: $\tau_E \sim \tau_E + \beta_{\rm Hw}$. In the Euclidean path integral formulation, we regard this periodicity as the inverse of the Hawking temperature of the black hole: $T_{\rm Hw} = \frac{1}{\beta_{\rm Hw}} = \frac{1}{2\pi\sqrt{k\alpha'}}$ because in the Matsubara formalism, the periodicity of the Euclidean time corresponds to the inverse temperature. Note that in the large $k$ semi-classical limit, we have vanishing Hawking temperature so that the back-reaction associated with the Hawking radiation is negligible and the black hole geometry is infinitely long-lived (i.e. eternal). As we mentioned in section \ref{sec:2-3}, it is also interesting to note that the Hawking temperature does not depend on the mass $m$ of the two-dimensional black hole. In the context of the dual LST defined in section \ref{sec:2-3}, we will regard this temperature as the (non-perturbative) Hagedorn temperature of the LST.
789: 
790: There are several derivations of the Hawking temperature other than the Euclidean path integral method. Recently \cite{Robinson:2005pd,Iso:2006wa} proposed a new derivation based on the gravitational anomaly in the vicinity of the event horizon.\footnote{See also \cite{Christensen:1977jc} for a derivation based on the trace anomaly.} We briefly review their derivation focusing on the two-dimensional case (see \cite{Iso:2006ut,Murata:2006pt,Vagenas:2006qb,Setare:2006hq} for related works). 
791: 
792: Let us take the very near-horizon limit (Rindler-limit)\footnote{Although there is nothing wrong with taking the Rindler-limit in the general relativity, the limit is a little bit subtle in the string theory because the string theory introduces a ``minimal size" (string scale) to the geometry. In our example, the central charge of the original two-dimensional black-hole and its very near horizon limit is different, so we need an extra compensation of the central charge in order to preserve the criticality condition. Since we are only interested in the classical thermodynamics, we will neglect this subtlety for a time being. See also the discussion of stretched horizon of the two-dimensional black holes in section \ref{sec:4}.} of the two-dimensional black hole:
793: \begin{align}
794: \dd s^2 = -\frac{2(r-r_0)}{\sqrt{k\alpha'}} \dd t^2 + \frac{\sqrt{k\alpha'}}{2(r-r_0)} \dd r^2 \ , \ \ \Phi = \mathrm{const} \ .
795: \end{align}
796: Now let us suppose we neglect the classically irrelevant in-falling modes of any scalar field propagating in the vicinity of the horizon $r=r_0$. The massless scalar fields with this boundary condition are effectively chiral, so it will show a gravitational anomaly:
797: \begin{align}
798: \nabla_\mu T^{\mu}_{\ \nu} = \frac{1}{\sqrt{-g}}\partial_\mu N^\mu_{\ \nu} 
799: \end{align}
800: with the explicit component expression
801: \begin{align}
802: N_{t}^t = N_{r}^r = 0 \ , \ \ N_t^r = \frac{1}{192\pi} \frac{4}{k\alpha'} \ , \ \ N_{r}^t = -\frac{1}{192\pi (r-r_0)^2}  \ . 
803: \end{align}
804: Especially, it shows a pure flux contribution
805: \begin{align}
806: \Phi = N_t^r|_{r=r_0} = \frac{1}{192\pi} \frac{4}{k\alpha'}  \ . \label{fl}
807: \end{align}
808: 
809: 
810: To cancel the gravitational anomaly, we need a quantum contribution that can be attributed to the Hawking radiation from the black hole. The black body radiation with the temperature $T_{\rm Hw}$ gives rise to the flux
811: \begin{align}
812: \Phi = \frac{\pi}{12}T^2_{\rm Hw} \ , \label{flt}
813: \end{align}
814: and the comparison between \eqref{fl} and \eqref{flt} establishes the Hawking temperature $T_{\rm Hw} = \frac{1}{2\pi\sqrt{k\alpha'}}$. We will later see similar effects related with the choice of boundary conditions of wavefunction at the horizon when we study D-brane motions in the two-dimensional black hole geometries.
815: 
816: Let us move on to the other thermodynamic quantities. Since the temperature does not depend on the mass of the two-dimensional black hole, we see that the black hole entropy is given by
817: \begin{align}
818: S(m) =  \beta_{\rm Hw} m = 2\pi\sqrt{\alpha' k}  m \ . \label{entr}
819: \end{align}
820: In higher dimensions, we could identify the entropy of the black hole as the area of the event horizon (i.e. the Bekenstein formula $S=\frac{A}{4\pi}$), but in the two-dimensional space-time, the event horizon is just a point and we cannot apply the Bekenstein formula. Instead, we have defined the entropy through the thermodynamic relation $\beta = \frac{\partial S}{\partial m} $. This formula predicts the high energy density of states in the LST. Assuming the microscopic explanation of the black hole entropy \eqref{entr} from the LST, the density of states of LST should be given by
821: \begin{align}
822: \rho(E) \sim e^{2\pi\sqrt{\alpha' k} E } \ ,
823: \end{align}
824: in the high energy limit $E\to \infty$.
825: \newpage
826: \sectiono{Two-dimensional Black Hole: CFT Viewpoint}\label{sec:3}
827: In this section, we review the two-dimensional black hole from the exactly solvable CFT viewpoint. We begin with the Euclidean version of the two-dimensional black hole  and then we move on to the Lorentzian two-dimensional black hole from an appropriate Wick rotation. This is because the Euclidean version is much better understood than the Lorentzian counterpart.
828: 
829: The organization of this section is as follows. In section \ref{sec:3-1}, We begin with the classical geometries for the $SL(2;\br)/U(1)$ coset model that yields an exact CFT model for the two-dimensional black hole system. In section \ref{sec:3-2}, we review the Euclidean spectrum of the $SL(2;\br)/U(1)$ coset model. In section \ref{sec:3-3}, we deal with the Lorentzian case in detail. In section \ref{sec:3-4}, we comment the duality between $SL(2;\br)/U(1)$ coset model and the $\mathcal{N}=2$ Liouville theory and discuss implications of the associated winding tachyon condensation.
830: 
831: \subsection{Classical geometries for $SL(2;\br)/U(1)$ coset}\label{sec:3-1}
832: From the world-sheet viewpoint, the reason why we are interested in the two-dimensional black hole system is that we can quantize the string theory on it by using the $SL(2;\br)/U(1)$ coset CFT. In this subsection, we would like to overview the correspondence between the $SL(2;\br)/U(1)$ coset model and the two-dimensional black hole system from the gauged Wess-Zumino-Novikov-Witten (WZNW) model construction \cite{Witten:1991yr,Elitzur:1991cb,Mandal:1991tz}.
833: 
834: It is possible to define the coset CFT such as $SL(2;\br)/U(1)$ model purely from the algebraic viewpoint (at least at the level of the left-right chiral $SL(2;\br)/U(1)$ representations: the most difficult point is to construct the modular invariant combinations), but we would like to begin with the Lagrangian construction based on \cite{Witten:1991yr}. This is because the construction directly gives the geometric interpretation of the model as the non-linear sigma model on the two-dimensional black hole in the semi-classical limit $(k \to \infty)$. The path integral formulation based on the Lagrangian  can also be used to derive the (formally) modular invariant partition function of the Euclidean two-dimensional black hole \cite{Hanany:2002ev} (see appendix \ref{part}).
835: 
836: The ungauged WZNW model for a general Lie group $G$ has the following action:
837: \begin{align}
838: S_{WZNW}(g) = \frac{\kappa}{8\pi} \int_{\Sigma} \dd^2 x \sqrt{|\gamma|}\gamma^{ij} \mathrm{tr} (g^{-1}\partial_i g g^{-1} \partial_j g) + i\kappa\Gamma(g) \ . \label{WZWa} 
839: \end{align}
840: The Wess-Zumino term $\Gamma(g)$ is given by 
841: \begin{align}
842: \Gamma(g) = \frac{1}{12\pi} \int_B \mathrm{tr} (g^{-1}\dd g\wedge g^{-1}\dd g\wedge g^{-1}\dd g ) \ ,
843: \end{align}
844: where $B$ is a three-dimensional manifold whose boundary is $\Sigma$. $\kappa$ denotes the level of the current algebra realized by the WZNW model.
845:  When the Lie group $G$ is compact, the level $\kappa$ should be quantized so that the Wess-Zumino term contributes to the path-integral uniquely with an arbitrary choice of $B$. In our case, however, since the Lie group is non-compact and $H^3(SL(2;\br),\br) = 0$, the quantization condition of the level $\kappa$ is not necessary.
846: 
847: Let $G$ be $SL(2;\br)$ for our discussion in the following. 
848: The action \eqref{WZWa} possesses a global $SL(2;\br)\times SL(2;\br)$ symmetry $g \to a gb^{-1}$, with $a,b \in SL(2;\br)$. Quantum mechanically, it will be elevated to a current algebra with the level $\kappa$: the chiral current
849: \begin{align}
850: j^A(z) = \kappa \mathrm{Tr}(T^A \partial g g^{-1}) \ ,
851: \end{align}
852: with $T^3= \frac{1}{2}\sigma_2$, $T^{\pm} = \pm \frac{1}{2}(\sigma_3 \pm i\sigma_1)$, satisfies the OPE of the affine $\widehat{SL(2;\br)_{\kappa}}$  current algebra
853: \begin{align}
854: \begin{cases} j^3(z)j^3(0) \sim -\frac{\kappa}{2z^2} \cr
855: j^3(z) j^{\pm}(0) \sim \pm \frac{1}{z}j^{\pm}(0) \cr
856: j^+(z)j^{-}(0) \sim \frac{\kappa}{z^2} - \frac{2}{z}j^{3}(0) 
857: \end{cases} \ .
858: \end{align}
859: The  bosonic $SL(2;\br)$ WZNW model has the central charge $c = \frac{3\kappa}{\kappa-2}$. 
860: We will gauge the anomaly free subgroup of the global symmetry of the $SL(2;\br)$ WZNW model to obtain the Lagrangian formulation for the coset CFT.
861: 
862: Let us first begin with the Euclidean coset. The $SL(2;\br)$ WZNW model has a negative-signature direction in $J^3 \sim i\sigma_2$ and the target space is a Lorentzian manifold. We gauge the (compact) $U(1)$ subgroup generated by
863: \begin{align}
864: \delta g = \epsilon (i\sigma_2 \cdot g + g \cdot (i\sigma_2)) \ , 
865: \end{align}
866: or by setting
867: \begin{align}
868: a = b^{-1} = h= \begin{pmatrix} \cos\epsilon & \sin\epsilon \\ -\sin\epsilon  & \cos\epsilon \end{pmatrix}  \ . \label{gauget}
869: \end{align}
870: 
871: To promote the global axial symmetry $g \to hgh $ to the gauge symmetry, we have to introduce the gauge connection $A_i$ that transforms as $\delta A_i = -\partial_i \epsilon$ under the gauge transformation \eqref{gauget} with a space varying gauge parameter $\epsilon(x_i)$. The covariantized action reads
872: \begin{align}
873: S_{\mathrm{gauged}} &= S_{\mathrm{WZNW}}(g) + \cr &+ \frac{\kappa}{2\pi} \int \dd^2z \bar{A} \mathrm{Tr}\left(i\sigma_2 g^{-1}\partial g\right) + A\mathrm{Tr}\left(i\sigma_2 \bar{\partial} g ^{-1}\right) + A\bar{A}\left(-2 + \mathrm{Tr}\left(i\sigma_2 g i\sigma_2 g^{-1}\right) \right) \ .  \label{gWZW}
874: \end{align}
875: We call this gauged WZNW model as $SL(2;\br)^{(A)}/U(1)$ axial coset.
876: To obtain the classical geometry of the $SL(2;\br)^{(A)}/U(1)$ coset CFT, we will integrate out the gauge field by fixing the gauge \footnote{In terms of the Euler angle parametrization (see appendix \ref{a-1-2}), $g = e^{i\sigma_2 \frac{t-\phi}{2}} e^{r\sigma_1}e^{i\sigma_2\frac{t+\phi}{2}}$ and we set $t=0$, where $\phi = \theta - \frac{\pi}{2}$. It is clear that this gauge fixing is always possible and unique.}
877: \begin{align}
878: g = \cosh r + \sinh r \begin{pmatrix} \cos\theta & \sin\theta \\ \sin\theta & -\cos\theta \end{pmatrix} \ .
879: \end{align}
880: The resulting sigma model for the gauge fixed coordinate $(r,\theta)$ is given by the action
881: \begin{align}
882: S = \frac{\kappa}{2\pi} \int \dd^2z \left(\partial r \bar{\partial} r + \tanh^2 r \partial \theta \bar{\partial} \theta \right) \ , \label{axcea}
883: \end{align}
884: with the dilaton gradient $e^{2\Phi} = \frac{k}{\mu \cosh^2 r}$, which originates from the one-loop determinant factor for the gauge field $A_i$. The geometry one can read from the sigma model action is the Euclidean two-dimensional black hole we have introduced in section \ref{sec:2}.
885: 
886: In the bosonic coset model, we expect a perturbative (and nonperturbative) $\alpha'$ corrections for this gauge fixing procedure and the corrected sigma model was proposed in \cite{Dijkgraaf:1992ba}. In the supersymmetric Kazama-Suzuki coset, it is believed that there is no perturbative $\alpha'$ corrections to the metric. Nonperturbative corrections which will be reviewed in section \ref{sec:3-4}, however, are present and they are one of the key elements to  understand the ``black hole - string transition".
887: 
888: The Lorentzian coset is obtained by gauging the non-compact subgroup
889: \begin{align}
890: \delta g = \epsilon\left(\sigma_3 g + g \sigma_3\right) \ .
891: \end{align}
892: We fix the gauge by setting
893: \begin{align}
894: g = \begin{pmatrix} a & u \\ -v & a \end{pmatrix}
895: \end{align}
896: with the determinant constraint $uv = 1- a^2$. The gauge fixing condition is valid for $1-uv>0$ and we can see that $u$ and $v$ are gauge invariant coordinates. With this gauge fixing condition,\footnote{Strictly speaking, the coset is a double cover of the $(u,v)$ plane, where the two-sheets are distinguished by the signature of $a$. We will neglect this small subtlety throughout the thesis.} the target space is spanned by the $(u,v)$ plane. After integrating out the gauge field $A_i$, the resulting sigma model is given by
897: \begin{align}
898: S = -\frac{\kappa}{4\pi} \int \dd^2x \sqrt{|\gamma|}\frac{\gamma^{ij}\partial_i u\partial_j v}{1-uv} \ ,
899: \end{align}
900: which reproduces the classical two-dimensional black hole system discussed in section \ref{sec:2}. Here we have used the Lorentzian signature world-sheet so that the sigma model with a Lorentzian  signature target space is well-defined.
901: 
902: In this thesis, we mainly focus on the supersymmetric generalization of the two-dimensional black hole based on the supersymmetric $SL(2;\br)_k/U(1)$ coset model. The starting point is the bosonic $SL(2;\br)_\kappa /U(1)$ coset mode with the level $\kappa = k+2$ bosonic current algebra. In addition to the bosonic action \eqref{gWZW}, we introduce the fermionic part:
903: \begin{align}
904: S_f = \frac{1}{2\pi} \int \dd^2z \left(\psi^+(\bar{\partial}-\bar{A})\psi^- + \psi^- (\bar{\partial}+\bar{A}) \psi^{+} + \tilde{\psi}^+ (\partial -A) \tilde{\psi}^-  + \tilde{\psi}^-(\partial +A)\tilde{\psi}^+ \right) \ , 
905: \end{align}
906: with the OPE $\psi^+(z)\psi^-(0) \sim 1/z$, $\psi^{\pm}(z)\psi^{\pm}(0) \sim 0$.
907: Let us concentrate on the Euclidean case for definiteness. From the Kazama-Suzuki construction, we can realize the $\mathcal{N}=2$ superconformal symmetry on the supersymmetric $SL(2;\br)/U(1)$ coset model. The explicit realization is given by
908: \begin{align}
909: T(z) & = \frac{1}{k}(j^{1}j^{1} +j^2j^2) - \frac{1}{2}(\psi^+\partial \psi^- - \partial \psi^+ \psi^-) \cr
910: G^{\pm}(z) &= \frac{1}{\sqrt{k}} \psi^{\pm} j^{\mp} \cr
911: J(z) & = \psi^+\psi^- + \frac{2}{k}(j^{3} + \psi^+\psi^-) \ ,
912: \end{align}
913: whose central charge is given by $c = 3(1+\frac{2}{k}) $.
914: The gauging current is defined by $J^3 = j^3 + \psi^+ \psi^-$, which commutes with all the elements of the $\mathcal{N}=2$ superalgebra. The fermionic part of the Lorentzian case is obtained from the analytic continuation by formally replacing $\psi^+ = \frac{1}{\sqrt{2}}(\psi^1 + i \psi^2) \to \frac{1}{\sqrt{2}}(\psi^1 + \psi^3)$ and $\psi^- = \frac{1}{\sqrt{2}}(\psi^1 - i \psi^2) \to \frac{1}{\sqrt{2}}(\psi^1 - \psi^3)$ .
915: 
916: From the path integral viewpoint, instead of treating the (Euclidean) $SL(2;\br)^{(A)}/U(1)$ coset, it is more convenient to study the equivalent description based on the $\mathbb{H}_3^+/\br$ coset model. The $\mathbb{H}_3^+ = SL(2;\mathbb{C})/SU(2)$ model is defined by the sigma model on the upper sheet
917: \begin{align}
918: g =\begin{pmatrix} a & u \\ \bar{u}  & b \end{pmatrix} =  \begin{pmatrix} e^{\phi} & e^{\phi}\bar{\gamma} \\ e^{\phi}\gamma  & e^{\phi}\gamma\bar{\gamma} + e^{-\phi} \end{pmatrix} \ ,
919: \end{align}
920: where we have introduced a real field $\phi$ and a complex field $\gamma$ with its complex conjugation $\bar{\gamma}$. The sigma model has the action
921: \begin{align}
922: S = -\frac{\kappa}{2\pi} \int \dd^2z \left(\partial\phi \bar{\partial}\phi + e^{2\phi}\partial\gamma\bar{\partial}\gamma \right)  \ .
923: \end{align}
924: The model has a positive definite action and the path integral is well-defined (unlike the ungauged $SL(2;\br)$ WZNW model).\footnote{However, the model has an imaginary $H_3$ flux, so the physical interpretation is unclear.} The two-dimensional Euclidean black hole is obtained by axially gauging the noncompact $U(1)$ direction $\sigma_2$ (e.g. one can choose a gauge $a = b$ and obtain the metric $ds^2 = \kappa\frac{dud\bar{u}}{1+|u|^2}$). This construction has the advantage that the parent sigma model has a definite Euclidean path integral while the $SL(2;\br)/U(1)$ coset does not because the parent WZNW model has a Lorentzian signature. 
925: 
926: So far, we have studied the axial coset of the $SL(2;\br)$ WZNW model, but it is possible to gauge the vector symmetry 
927: \begin{align}
928: \delta g = \epsilon (i\sigma_2 \cdot g - g \cdot (i\sigma_2)) \ .
929: \end{align}
930: This symmetry has a fixed point, and the corresponding effective action
931: \begin{align}
932: S = \frac{\kappa}{2\pi} \int \dd^2z \left(\partial \rho \bar{\partial} \rho + \frac{1}{\tanh^2 \rho} \partial \tilde{\theta} \bar{\partial} \tilde{\theta} \right) \ , \label{efftru}
933: \end{align}
934: which is also known as the trumpet model, has a singularity at $\rho=0$. However, from the algebraic coset viewpoint, there is no singularity at all. We have just replaced right moving $\bar{J}_3$ current of the $SL(2;\br)$ WZNW model with $-\bar{J}_3$.
935: 
936: In order to specify the model, we have to determine the periodicity of the variable $\tilde{\theta}$ in the effective action \eqref{efftru}. The angular variable $\theta$ in the cigar geometry has a natural periodicity $2\pi$ coming from the $SL(2;\br)$ WZNW model. In the trumpet case, there is no apriori natural periodicity of $\tilde{\theta}$ because it is non-contractible loop in $SL(2;\br)$ and any periodicity is allowed if we study the universal cover of the $SL(2;\br)$. In terms of the Euler angle parametrization (see appendix \ref{a-1-2}), $g = e^{i\sigma_2 \frac{t-\phi}{2}} e^{r\sigma_1}e^{i\sigma_2\frac{t+\phi}{2}}$ and we set $\phi=0$. In this sense, the natural periodicity for $t = \tilde{\theta}$ is $2\pi$. We {\it define} that the $SL(2;\br)^{(V)}/U(1)$ vector coset model has $2\pi$ periodicity in $\tilde{\theta}$.
937: 
938: From the algebraic construction, the vector coset merely changes the sign convention of the right-moving current $\bar{J}_3$, which reminds us of the T-duality or mirror symmetry. Along this line of reasoning, an alternatively good definition of the vector coset is to take $ 2\pi/k$ periodicity in $\tilde{\theta}$.\footnote{This convention is the one given in \cite{Dijkgraaf:1992ba}.} This is indeed motivated by Bucsher's T-duality rule: if we perform the T-duality to our original cigar model \eqref{axcea} with $2\pi$ periodicity in $\theta$, we obtain the trumpet model with $ 2\pi/k$ periodicity. We call this model as ``$\bz_k$ orbifold of the vector coset $SL(2;\br)^{(V)}/U(1)$". The $\bz_k$ orbifold of the $SL(2;\br)^{(V)}/U(1)$ or $\bz_k$ orbifold of the trumpet model is same as the cigar model as a CFT (up to a GSO projection in the supersymmetric case).\footnote{Let us mention the similar structure in the $SU(2)$ case. It is known that the axial coset $SU(2)^{(A)}/U(1)$ is the $\bz_k$ orbifold of the vector coset $SU(2)^{(V)}/U(1)$. However, in this case, $\bz_k$ orbifold of the vector coset $SU(2)^{(V)}/U(1)$ is the same model as the original $SU(2)^{(V)}/U(1)$. Therefore, we can also say that the T-dual of the $SU(2)^{(A)}/U(1)$ coset is the $SU(2)^{(V)}/U(1)$. In the $SL(2;\br)$ case, although the asymptotic spectrum of the $\bz_k$ orbifold of the $SL(2;\br)^{(V)}/U(1)$ coincides with that of the $SL(2;\br)^{(V)}/U(1)$, the (un-regularized) partition function is different from each other. Here we implicitly assumed that $k$ is an integer, but the situation is more involved when $k$ is not an integer because the meaning of the $\bz_k$ orbifold is obscure. Note that we have defined the $\bz_k$ orbifold of the $SL(2;\br)^{(V)}/U(1)$ as the T-dual of the $SL(2;\br)^{(A)}/U(1)$, which perfectly makes sense even for irrational $k$.}
939: 
940: Since the vector coset model with the trumpet geometry is related to the T-duality of the cigar geometry, there should be no singularity at all in the vector coset model as a CFT. What happens to the apparent singularity of the classical geometry? We will come back to this problem in section \ref{sec:3-4}.
941: 
942: The equivalence between the axial coset and the ($\bz_k$ orbifold of) vector coset leads to a remarkable observation made in \cite{Dijkgraaf:1992ba} --- the duality between the singularity and the horizon. If we gauge the vector symmetry for the Lorentzian coset, we end up with the same Lorentzian two-dimensional black hole.\footnote{This is up to global duplications. For instance, if one considers an axial coset of the universal cover of the $SL(2;\br)/U(1)$ coset, we have an infinite copies of the Lorentzian two-dimensional black holes.} However, the analytic continuation of the trumpet geometry (vector coset) has a natural interpretation as the region inside the singularity:
943: \begin{align}
944: \dd s^2 = \alpha' k\left(- \frac{1}{\tanh^2\rho} \dd t^2 + \dd\rho^2\right) \ .
945: \end{align}
946: 
947: In this way, the axial-vector duality suggests the duality of the region parametrized $uv$ and $1-uv$ while keeping $t$. In particular, it exchanges the region outside the horizon and the one inside the singularity. This duality would shed a new light on the physics of the black hole and especially, relation between the T-duality and ``winding tachyon" in the Lorentzian two-dimensional black hole. It is, however, fair to say that the precise physical meaning of the duality is far from being well-understood in the Lorentzian signature.
948: 
949: To close this section, we discuss the pure two-dimensional background by setting $k=1/2$ or $\kappa = 9/4$. Then the model is a critical string theory by itself and one can regard it as a pure two-dimensional background \cite{Witten:1991yr}. In the Euclidean signature, there is a proposed dual matrix model \cite{Kazakov:2000pm} and the theory is supposedly exactly solvable. In the context of more general dilaton gravity in the two-dimensional space-time, the classical solutions are investigated in \cite{Grumiller:2005sq}.
950: 
951: 
952: \subsection{Euclidean spectrum}\label{sec:3-2}
953: Let us study the spectrum of the Euclidean $SL(2;\br)/U(1)$ coset model. 
954: 
955: \subsubsection{algebraic coset}\label{sec:3-2-1}
956: We begin with the algebraic structure of the coset model from the noncompact para-fermion construction. 
957: 
958: Let $\Phi_{jm}(z)$ be holomorphic part of the primary fields of $SL(2;\br)$ WZNW model with bosonic level $\kappa$. It has a (bosonic) left-moving $j^3_0$ eigenvalue $m$:
959: \begin{align}
960: j^3(z) \Phi_{jm}(0) = m \frac{\Phi_{jm}}{z} \ .
961: \end{align}
962:  It is also a holomorphic part of the primary for the supersymmetric $SL(2;\br)$ WZNW model with the same eigenvalue for $J^3_0$. Here, the supersymmetric current $J^3$ is defined by
963: \begin{align}
964: J^3 = j^3 -\psi^-\psi^+ \ .
965: \end{align}
966: For later convenience, we introduce the bosonized current\footnote{Note that the gauging current must be time-like in the Euclidean coset.}
967: \begin{align}
968: \partial H &= i\psi^-\psi^+ \cr
969: J^3 &= -\sqrt{\frac{k}{2}}\partial X_3  \cr
970: j^3 &= J^3 + i\partial H = -\sqrt{\frac{\kappa}{2}}\partial x_3 \ .
971: \end{align}
972: Using $X_3$, or $x_3$, we can decompose $\Phi_{jm}$ as
973: \begin{align}
974: \Phi_{jm} = U_{jm} e^{m\sqrt{\frac{2}{k}}X_3} = V_{jm} e^{m\sqrt{\frac{2}{\kappa}}x_3} \ . \label{decb}
975: \end{align}
976: The para-fermion fields $U_{jm}$ and $V_{jm}$ have the conformal dimension
977: \begin{align}
978: \Delta(U_{jm}) &= \frac{-j(j+1)+m^2}{k} \cr
979: \Delta(V_{jm}) &= -\frac{j(j+1)}{\kappa-2} + \frac{m^2}{\kappa} \ , 
980: \end{align}
981: and they are (holomorphic) primaries of the supersymmetric Euclidean coset $SL(2;\br)/U(1)$ and the bosonic coset respectively. 
982: 
983: For the supersymmetric case, we should also decompose the $U(1)$ fermion current as 
984: \begin{align}
985: e^{inH} = e^{n\sqrt{\frac{2}{k}}X_3} e^{i\sqrt{\frac{c}{3}} X_R} \ , \label{decf}
986: \end{align}
987: where the bosonized $U(1)_R$-current (of the $\mathcal{N}=2$ SUSY algebra: see appendix \ref{SCA2}) is defined by
988: \begin{align}
989: J = i\sqrt{\frac{c}{3}}\partial X_R \ ,
990: \end{align}
991: where
992: \begin{align}
993: iH = \sqrt{\frac{2}{k}}X_3 + i\sqrt{1+\frac{2}{k}}X_R \ .
994: \end{align}
995: 
996: Under the decomposition \eqref{decb} and \eqref{decf}, (holomorphic) primary fields of the supersymmetric $SL(2;\br)/U(1)$ coset takes the form
997: \begin{align}
998: V^{n}_{jm} = V_{jm} e^{i(\frac{2m}{k+2} +n)\sqrt{\frac{c}{3}}X_R} \ ,
999: \end{align}
1000: which has the conformal dimension
1001: \begin{align}
1002: \Delta(V_{jm}^n) = -\frac{j(j+1) + (m+n)^2}{k} + \frac{n^2}{2} \ ,
1003: \end{align}
1004: and the $U(1)_R$ charge
1005: \begin{align}
1006: R(V_{jm}^n) = \frac{2m}{k} + \frac{nc}{3} \ .
1007: \end{align}
1008: The (half) integer $n$ is the amount of the spectral flow of the $\mathcal{N} = 2 $ superconformal algebra. The structure of the descendants, depending on quantum numbers $(j,m,n)$, are completely fixed from that of the $SL(2;\br)$ (see appendix \ref{a-1}), or alternatively from the representation of the $\mathcal{N}=2 $ superconformal algebra.
1009: 
1010: \subsubsection{spectrum from partition function}\label{sec:3-2-2}
1011: 
1012: To obtain the full spectrum of the CFT from the holomorphic data discussed in section \ref{sec:3-2-1}, we need to combine left-moving parts and right-moving parts in a consistent way. For example, in the compact $SU(2)$ WZNW model, the complete classification of the modular invariant partition function (and hence the spectrum) is given by the so-called ADE classification. In the non-compact case, we have not yet achieved such a systematic classification. Physically, however, we are primarily interested in the two-dimensional black hole interpretation of the $SL(2;\br)/U(1)$ coset model, so we will only consider the simplest realization from the gauged WZNW model as we have focused in section \ref{sec:3-1}.
1013: 
1014: We begin with the partition function for the bosonic Euclidean two-dimensional black hole \cite{Hanany:2002ev}
1015: \begin{align}
1016: Z_{\mathbb{H}^{3(A)}_{+}/\br} = \int_{\Sigma} \frac{\dd u^2}{\tau_2} \frac{e^{\frac{u_2^2}{\tau_2}}}{\sqrt{\tau_2}|\theta_1(\tau,u)|^2 } \sqrt{\tau_2}|\eta(\tau)|^2 \sum_{m,\omega\in \bz} e^{-\frac{\pi \kappa}{\tau_2}|\omega\tau - m + u|^2} \ . \label{ppo}
1017: \end{align}
1018: See appendix \ref{part} for a summary of various partition functions. Unfortunately, the partition function is divergent, and the leading divergence comes from the integration near $u_1 = u_2 = 0$. The diverging factor could be attributed to the volume divergence in the radial $\rho$ direction of the cigar model. Thus, at the leading order, we have
1019: \begin{align}
1020: Z_{\mathbb{H}^{3(A)}_{+/\br}} \sim \frac{1}{2\pi} (\log\epsilon) Z_{\mathrm{free}}(\tau) Z_{\sqrt{\kappa}}(\tau) + \text{finite part} \ , \label{lpf}
1021: \end{align}
1022: which gives the asymptotic degrees of freedom realized by a free non-compact boson (with the linear dilaton):
1023: \begin{align}
1024: Z_{\mathrm {free}}(\tau) = \frac{1}{\sqrt{\tau_2}|\eta(\tau)|^2} \ ,
1025: \end{align}
1026:  and a compact boson with radius $R^2 = \kappa$. 
1027: \begin{align}
1028: Z_{\sqrt{\kappa}} = \frac{\sqrt{\kappa}}{\sqrt{\tau_2}|\eta(\tau)|^2} \sum_{m,\omega\in \bz} e^{-\frac{\pi \kappa}{\tau_2}|\omega\tau - m|^2} \ .
1029: \end{align}
1030: The appearance of the compact boson is due to the summation over the lattice $(n,\omega)$, which arises from the zeros of $\theta_1(\tau,u)$ in \eqref{ppo}.
1031: 
1032: A more precise (but formal) manipulation \cite{Hanany:2002ev} leads to the following decomposition of the partition function:
1033: \begin{align}
1034: Z_{\mathbb{H}^{3(A)}_{+/\br}} = \int^{-1/2}_{-(\kappa-1)/2} \dd j \mathrm{Tr}_{\hat{D}^+_j\otimes\hat{D}^+_j} q^{L_0}q^{\bar{L}_0} + \sum_{\omega,n}\int_0^\infty \dd p 2\rho(p) \mathrm{Tr}_{\hat{C}_{-\frac{1}{2}+ip}\otimes\hat{C}_{-\frac{1}{2}+ip}} q^{L_0}q^{\bar{L}_0} + \dots \ , \label{decompp}
1035: \end{align}
1036: where the Hilbert space $\hat{D}^+_j\otimes\hat{D}^+_j$ is the discrete representations of the $SL(2;\br)$ with the constraints $J_0^3 -\bar{J}_0^3 = n$, $J_0^3 + \bar{J}_0^3 = \kappa \omega $ and no contribution from the $J^3_{n<0}$ oscillators.
1037: The same restriction is imposed on the continuous representations $\hat{C}_{-\frac{1}{2}+ip}\otimes\hat{C}_{-\frac{1}{2}+ip}$. The density of states for the continuous representations is given by
1038: \begin{align}
1039: \rho(p) = \frac{1}{2\pi} 2 \log \epsilon + \frac{1}{2\pi i} \frac{\partial}{2\partial p} \log \frac{\Gamma(-ip+\frac{1}{2}-m)\Gamma(-ip+\frac{1}{2}+\bar{m})}{\Gamma(+ip+\frac{1}{2}+\bar{m})\Gamma(+ip+\frac{1}{2}-m)} \ , \label{dnsi}
1040: \end{align}
1041: where $m = \frac{1}{2}(n+\kappa \omega)$, $\bar{m} = -\frac{1}{2}(n-\kappa\omega)$ are the eigenvalues of $J_0^3$ and $\bar{J}_0^3$. The density of states appearing here is consistent with the reflection amplitude (or sphere two-point function) of the two-dimensional black hole as we will review in section \ref{sec:3-2-3}. The leading diverging part proportional to $\log \epsilon$ agrees with \eqref{lpf}. 
1042: 
1043: There are, however, several subtleties associated with the decomposition \eqref{decompp}. First of all, the expression \eqref{decompp} is {\it not} modular invariant although our starting point \eqref{ppo} is formally invariant. The failure is due to the nontrivial $p$ dependent density of states \eqref{dnsi} and the contribution from the discrete series. In other words, the regularization rule for the character decomposition \eqref{decompp} does not preserve the modular invariance. The only one can say is that the leading part, (or the partition function per unit volume as $\epsilon \to 0$) is modular invariant \eqref{lpf}. Another subtlety is related to the omitted terms in  \eqref{decompp}. The regularization procedure proposed in \cite{Hanany:2002ev} (see also \cite{Maldacena:2000hw} for related models) actually leaves us with finite terms that could not be written as the character appearing in \eqref{decompp} \cite{Israel:2004ir}. Again this depends on the regularization scheme and one natural (but not unique) solution is to omit this part as we will implicitly assume in the following.
1044: 
1045: Despite all these subtleties, the decomposition \eqref{decompp} seems to capture important physics of the two-dimensional black hole. In particular, it predicts the existence of the discrete spectrum localized near the tip of the cigar. Indeed the range of the discrete representations $\frac{-\kappa+1}{2}<j<-\frac{1}{2}$ will be independently checked by the Cardy analysis of the boundary states for $SL(2;\br)/U(1)$ coset model as we will see in section \ref{sec:6-2-3}. Also, the minisuperspace analysis for the two-dimensional black hole reproduces the zero-slope limit of the results given here including the density of states \eqref{dnsi}. We will review the mini-superspace analysis in section \ref{sec:3-2-3}.
1046: 
1047: Before moving on to the mini-superspace analysis, we will briefly present a generalization to the supersymmetric $SL(2;\br)/U(1)$ coset model. The partition function is given by
1048: \begin{align}
1049: Z^{(NS)}(\tau) = \int_{\Sigma} \frac{\dd u^2}{\tau_2} \frac{|\theta_3(\tau,u)|^2}{\sqrt{\tau_2}|\theta_1(\tau,u)|^2 } \sqrt{\tau_2}|\eta(\tau)|^2 \sum_{m,\omega\in \bz} e^{-\frac{\pi k}{\tau_2}|\omega\tau - m + u|^2} \ 
1050: \end{align}
1051: and the decomposition to the character is obtained as 
1052: \begin{align}
1053: \int^{-1/2}_{-(k+1)/2} \dd j \mathrm{Tr}_{\hat{D}^+_j\otimes\hat{D}^+_j} q^{L_0}q^{\bar{L}_0} + \sum_{\omega,n \in \bz}\int_0^\infty \dd p 2 \rho(p) \mathrm{Tr}_{\hat{C}_{-\frac{1}{2}+ip}\otimes\hat{C}_{-\frac{1}{2}+ip}} q^{L_0}q^{\bar{L}_0} + \dots \ , \label{decomps} \ .
1054: \end{align}
1055: In this case, the trace should be taken over the (NS-NS) Hilbert space of the supersymmetric coset instead of the bosonic one. Explicitly
1056: \begin{align}
1057: \mathrm{Tr}_{\hat{C}_{-\frac{1}{2}+ip}\otimes\hat{C}_{-\frac{1}{2}+ip}} q^{L_0}q^{\bar{L}_0} = q^{\frac{p^2+m^2}{k}}\bar{q}^{\frac{p^2+\bar{m}^2}{k}} \frac{|\theta_3(\tau)|^2}{|\eta(\tau)|^6} \ , 
1058: \end{align}
1059: and 
1060: \begin{align}
1061:  \int^{-1/2}_{-(k+1)/2} \dd j \mathrm{Tr}_{\hat{D}^+_j\otimes\hat{D}^+_j} q^{L_0}q^{\bar{L}_0} &=  \sum_{\omega,n \in {\mathbf{Z}}} \sum_{j \in \mathcal{J}_{\omega,n}} \chi_{\mathrm{dis}, 1+j+\frac{k}{2}, m +\frac{k}{2}}(\tau) \chi_{\mathrm{dis}, 1+j+\frac{k}{2}, \bar{m} + \frac{k}{2}}(\bar{\tau}) \cr
1062: \mathcal{J}_{\omega,n} &= \left.\left[-\frac{k+1}{2},-\frac{1}{2}\right.\right) \cap \left(\frac{k\omega-n}{2} + \mathbb{Z}\right) \cr
1063: \chi_{\mathrm{dis}, j,j+n}({\tau}) &= \frac{q^{\frac{(j+n)^2}{k}-\frac{1}{4k}}}{1+q^{n+1/2}}\frac{\theta_3(\tau,0)}{\eta(\tau)^3} \ .
1064: \end{align}
1065: The partition functions of the other sectors are readily obtained by performing the spectral flow symmetry of the $\mathcal{N}=2$ SCA. We note that in order to obtain a superstring compactification with other sectors (such as $\mathcal{N}=2$ minimal models), we have to project down to the sectors with integral $U(1)_R$ -charge so that the space-time supersymmetry is well-defined (GSO projection).
1066: 
1067: 
1068: We can read some important physics from the spectrum of the Euclidean two-dimensional black hole:
1069: \begin{itemize}
1070: 	\item The continuous representations have a mass gap, which is consistent with the (asymptotic) linear dilaton background. Due to the mass gap, would-be graviton is massive, which is again consistent with the statement that the LST is non-gravitational theory.
1071: 	\item The discrete representations correspond to local dynamical degrees of freedom that has a winding quantum number being localized near the tip of the cigar. From the space-time point of view, they are normalizable deformation of the background localized in the vicinity of the singularity. The improved unitarity bound perfectly agrees with the geometrical normalizability condition discussed in section \ref{sec:2-2-2}.
1072: \end{itemize}
1073: 
1074: The improved unitarity bound has an important application to obtain the normalizable deformations of the LST.
1075: As promised, we will derive the geometrical bound \eqref{normc} from the improved unitarity bound of the $SL(2;\br)/U(1)$ coset model. The dual string theory for the generalized conifold
1076: \begin{align}
1077: z_1^{n} + z_2^2 + z_3^2 + z_4^2 = 0 \ 
1078: \end{align}
1079: is given by the $(n-2)$-th $\mathcal{N}=2$ minimal model coupled with $SL(2;R)/U(1)$ coset with the level $k = \frac{2n}{n+2}$. 
1080: 
1081: The vertex operators corresponding to massless deformations of the geometry can be obtained by combining (anti-)chiral primary operators of the $\mathcal{N}=2$ minimal model and the $SL(2;\br)/U(1)$ coset model restricted to $h=\bar{h}=\frac{1}{2}$. Labeling the chiral primaries of the minimal model by $l$ ($0 \le l \le n-2)$ with the $U(1)_R$ charge $Q_R = \frac{l}{n}$, we obtain the conformal condition:
1082: \begin{align}
1083: \frac{l}{n} + \frac{2m}{k} = 1 \ .
1084: \end{align}
1085: 
1086: On the other hand, the improved unitarity constraint is 
1087: \begin{align}
1088: 1 \le 2m \le 1+k \ ,
1089: \end{align}
1090: which gives the constraint
1091: \begin{align}
1092: l = 0 , 1 , \cdots , \left[\frac{n-2}{2}\right] \ .
1093: \end{align}
1094: The bound is in perfect agreement with \eqref{normc}.
1095: 
1096: 
1097: \subsubsection{minisuperspace analysis}\label{sec:3-2-3}
1098: For a complementary method to read the spectrum of the sigma model is to use the point particle approximation known as the mini-superspace approximation.
1099: 
1100: Let us consider the Euclidean two-dimensional black hole background, known
1101: as `cigar geometry':
1102: \begin{equation}
1103: \dd s^2 \equiv G_{ij} \dd x^i \dd x^j = 2k (\dd \rho^2 + \tanh^2\rho
1104: \dd \theta^2) \qquad \mbox{and} \qquad e^{\Phi} =
1105: \frac{e^{\Phi_0}}{\cosh\rho} ~. \label{Euclidean cigar}
1106: \end{equation}
1107: Recall that $k$ sets characteristic curvature radius in unit of the
1108: string scale and hence string world-sheet effects, while $e^{\Phi_0}$
1109: sets the maximum value of the string coupling at the tip $\rho=0$ of
1110: the cigar geometry. We shall assume the limit $k \gg 1$ and
1111: $e^{-\Phi_0} \gg 1$: this limit suppresses both string world-sheet
1112: and space-time quantum effects and facilitates to truncate closed
1113: string spectrum to zero-modes, viz. to mini-superspace approximation.
1114: 
1115: In the mini-superspace approach, difference between bosonic strings
1116: (with no world-sheet supersymmetry) and fermionic strings (with
1117: ${\cal N}=2$ world-sheet supersymmetry) becomes unimportant. The
1118: closed string Hamiltonian $L_0 + \overline{L}_0$ is reduced in the
1119: mini-superspace approximation to the target space Laplacian
1120: $\Delta_0$, where:
1121: \begin{align}
1122: \Delta_0 &= \frac{1}{e^{-2\Phi} \sqrt{G}} \partial_i
1123: \left(e^{-2\Phi} \sqrt{G} G^{ij} \partial_j\right) \equiv
1124: -\frac{1}{2k}[\partial_\rho^2 +2\coth2\rho\partial_\rho +
1125: \coth^2\rho\partial_\theta^2] ~. \label{Laplacian}
1126: \end{align}
1127: The Hamiltonian is defined with respect to the volume element:
1128: \begin{align}
1129: \dd \mbox{Vol} = e^{-2\Phi}\sqrt{G} \dd \rho \dd \theta := {2k}
1130: \sinh \rho \cosh \rho \dd \rho \, \dd \theta \equiv k \sinh 2\rho
1131: \dd \rho \, \dd \theta~, \label{vol cigar}
1132: \end{align}
1133: inherited from the Haar measure on the $SL(2;\br)$ group manifold.
1134: In the volume element, the dilaton factor $e^{-2\Phi}$ is taken into
1135: account, as the inner product for closed string states is defined by
1136: the world-sheet two-point correlators on the sphere. The normalized
1137: eigenfunctions are obtained straightforwardly \cite{Dijkgraaf:1992ba,Ribault:2003ss}. They
1138: are:
1139: \begin{align}
1140:  \phi_n^j(\rho,\theta) & =  -\frac{\Gamma^2(-j+\frac{|n|}{2})}{\Gamma(|n|+1)\Gamma(-2j-1)}
1141: e^{in\theta} \times\cr
1142: & \times \left[ \sinh^{|n|}\rho \cdot
1143: F\left(j+1+\frac{|n|}{2},-j+\frac{|n|}{2};|n|+1;-\sinh^2\rho
1144: \right)\right] ~ , \label{ef}
1145: \end{align}
1146: where $F(\alpha, \beta; \gamma; z)$ is the Gaussian
1147: hypergeometric function. These eigenfunctions correspond to the
1148: primary state vertex operators of conformal weights
1149: \begin{align}
1150: && h= \bar{h}= -\frac{j(j+1)}{k-2} + \frac{n^2}{4k} \qquad
1151: \mbox{or} \qquad h= \bar{h}= -\frac{j(j+1)}{k} + \frac{n^2}{4k}
1152: \end{align}
1153: for bosonic\footnote{The eigenvalue is actually proportional to
1154: $-\frac{j(j+1)}{k}+ \frac{n^2}{4k}$. We will return to this small mismatch at the end of this subsection.} and fermionic strings,
1155: respectively. We shall focus on the continuous series, parametrise
1156: the radial quantum number $j$ as $j= -\frac{1}{2}+ i\frac{p}{2}$
1157: $(p\in \br)$, and label the eigenfunctions as
1158: $\phi^p_n(\rho,\theta)$ instead of $\phi^j_n(\rho,\theta) $. We adopt
1159: the convention that, in the asymptotic region $\rho \sim \infty$,
1160: the vertex operators with $p>0$ corresponds to the incoming waves
1161: and those with $p<0$ corresponds to the outgoing waves. The
1162: eigenfunctions \eqref{ef} are then normalized as
1163: \begin{align}
1164: \Big(\phi^p_n, \phi^{p'}_{n'} \Big) = \delta_{n,n'} \Big\lb 2 \pi
1165: \delta(p-p')+ \cR_0(p',n) \, 2 \pi \delta(p+p') \Big\rb
1166: ~,\label{inner product}
1167: \end{align}
1168: where the inner product is defined with respect to the volume
1169: element \eqref{vol cigar}. Here, $ \cR_0(p,n)$ refers to the
1170: reflection amplitude of the mini-superspace analysis:
1171: \begin{align}
1172: \cR_0(p,n) =
1173: \frac{\Gamma(+ip)\Gamma^2(\frac{1}{2}-\frac{ip}{2}+\frac{n}{2})}
1174: {\Gamma(-ip)\Gamma^2(\frac{1}{2}+\frac{ip}{2}+\frac{n}{2})} \ .
1175: \label{cref amp}
1176: \end{align}
1177: That is, from the definition \eqref{ef}, the reflection amplitude is
1178: seen to obey the mini-superspace reflection relation:
1179: \begin{align}
1180: \phi^{-p}_n(\rho,\theta) = \cR_0(-p,|n|) \,
1181: \phi^{+p}_n(\rho,\theta)~. \label{cref rel}
1182: \end{align}
1183: We shall refer $\cR_0(p,n)$ as `mini-superspace' reflection
1184: amplitude, valid strictly within mini-superspace approximation at $k
1185: \rightarrow \infty$, and anticipate string world-sheet effects at
1186: finite $k$. Notice that no winding states wrapping around
1187: $\theta$-direction are present since by definition the
1188: mini-superspace approximation retains states with zero winding only.
1189: 
1190: Utilizing the analytic continuation formula of the hypergeometric
1191: functions:
1192: %
1193: \begin{align}
1194: F(\alpha,\beta;\gamma;z) &=
1195: \frac{\Gamma(\gamma)\Gamma(\beta-\alpha)}
1196: {\Gamma(\beta)\Gamma(\gamma-\alpha)}
1197: (-z)^{-\alpha}F(\alpha,\alpha+1-\gamma;
1198: \alpha+1-\beta;1/z) \cr
1199: &+ \frac{\Gamma(\gamma)\Gamma(\alpha-\beta)}
1200: {\Gamma(\alpha)\Gamma(\gamma-\beta)}(-z)^{-\beta}
1201: F(\beta,\beta+1-\gamma;\beta+1-\alpha;1/z)
1202: \label{eq:inv} \ ,
1203: \end{align}
1204: %
1205: the eigenfunction \eqref{ef} can be decomposed into
1206: %
1207: \begin{align}
1208: \phi^p_n(\rho,\theta) = \phi^p_{L,n}(\rho,\theta) + \cR_0(p,|n|)
1209: \phi^p_{R,n}(\rho,\theta) ~, \label{decomp ef} \end{align}
1210: %
1211: where
1212: %
1213: \begin{align}
1214: \phi^p_{L,n}(\rho,\theta) &\equiv e^{in\theta} (\sinh
1215: \rho)^{-1-ip}\,
1216: F\Big(\frac{1}{2}+\frac{ip+n}{2},\frac{1}{2}+\frac{ip-n}{2}; 1+ip;
1217: -\frac{1}{\sinh^2\rho} \Big) ~,\nn & \sim 
1218: e^{-\rho}e^{-ip\rho+in\theta} \qquad \mbox{at} \qquad \rho \,
1219: \rightarrow\, +\infty~ \label{phiL} \end{align}
1220: %
1221: and
1222: %
1223: \begin{align}
1224: \phi^p_{R,n}(\rho,\theta) &\equiv e^{in\theta} (\sinh
1225: \rho)^{-1+ip}\,
1226: F\Big(\frac{1}{2}-\frac{ip+n}{2},\frac{1}{2}-\frac{ip-n}{2}; 1-ip;
1227: -\frac{1}{\sinh^2\rho} \Big) \nn & \sim 
1228: e^{-\rho}e^{ip\rho+in\theta} \qquad \mbox{at} \qquad
1229:    \rho \, \rightarrow\, +\infty
1230: \label{phiR}
1231: \end{align}
1232: refer to the left- and the right-movers, respectively, at $\rho
1233: \rightarrow +\infty$, and $\cR_0(p, |n|)$ is defined in \eqref{cref
1234: amp}. Obviously, they are related to each other under the reflection of
1235: radial momentum: $\phi^{+p}_{R, n} = \phi^{-p}_{L, n}$, which is
1236: also evident from \eqref{decomp ef} and \eqref{cref amp}. These
1237: mini-superspace wave functions \eqref{decomp ef} constitute the
1238: starting point of constructing boundary states of D-brane in the
1239: Euclidean two-dimensional black hole background.
1240: 
1241: We close the mini-superspace analysis with remarks concerning Wick
1242: rotation of the results to the Lorentzian background and string
1243: world-sheet effects present at finite $k$.
1244: \begin{enumerate}
1245:  \item The decomposition of $\phi^p_n$ into $\phi^p_{L,n}$ and $\phi^p_{R,n}$ cannot
1246:  globally defined over the entire cigar geometry. They
1247:  are ill-defined around the tip $\rho =0$, and the reflection relation
1248: \eqref{cref rel} implies that $\phi^{-p}_n$ is not independent of
1249: $\phi^{+p}_n$. Therefore, of the continuous series, only the
1250: eigenfunctions $\phi^p_n$ with $p>0,~ n\in \bz$ span the physical
1251: Hilbert space of the closed strings on the Euclidean two-dimensional
1252: black hole. On the other hand, the situation will become further
1253: complicated once Wick rotated to the Lorentzian two-dimensional
1254: black hole.
1255: %%%
1256:  \item Notice that $\phi^p_n$ is not analytic
1257: with respect to the angular quantum number $n$ as it depends on its
1258: absolute value, $|n|$. This leads to the ambiguity for Wick rotation
1259: from Euclidean to Lorentzian background, under which roughly
1260: speaking $i n$ is replaced by energy $\omega$. As for the
1261: mini-superspace reflection amplitude $\cR_0(p, n)$, since
1262: $\cR_0(p,-n)= \cR_0(p,n)$ holds for all $n \in \bz$, it is
1263: unnecessary to take absolute value $|n|$ in \eqref{cref rel},
1264: \eqref{decomp ef}. When taking Wick rotation, we will start from the
1265: expression $\cR_0(p,|n|)$. In other words, we analytically continue
1266: $\cR_0(p,n)$ if $n > 0$ and $\cR_0(p,-n)$ if $n<0$.
1267: %%%
1268:  \item It is evident that $|\cR_0(p,n)|=1$, viz, the mini-superspace
1269:  reflection
1270:  amplitude is purely a phase shift in the Euclidean black hole
1271:  background. It is of
1272:  utmost importance that, in the Lorentzian black hole background,
1273:  $n$ is analytically continued to pure imaginary value,
1274:  and the modulus of the reflection amplitude becomes less than unity.
1275: %%%
1276:  \item For the fermionic Euclidean $SL(2; \br)/U(1)$ conformal field
1277:  theory, exact result for the reflection amplitude (i.e. taking account
1278:  of all string world-sheet effects) is known
1279:  \cite{Teschner:1997ft,Teschner:1999ug,Giveon:1999px,Giveon:1999tq}. In our notations, it is
1280: \begin{align}
1281: \cR(j,m,\bar{m}) = \nu(k)^{-2j-1}\,
1282: \frac{\Gamma(1+\frac{2j+1}{k})}{\Gamma(1-\frac{2j+1}{k})}
1283: \frac{\Gamma(2j+1)\Gamma(-j+m)\Gamma(-j-\bar{m})}{\Gamma(-2j-1)
1284: \Gamma(j+1+m)\Gamma(j+1-\bar{m})}, \label{qref amp}\end{align}
1285: %
1286: where
1287: %
1288: \begin{align} \nu(k)\equiv \frac{1}{\pi}\frac{\Gamma(1-\frac{1}{k})}
1289: {\Gamma(1+\frac{1}{k})}~, \qquad m=\frac{kw+n}{2}~, \qquad \bar{m} =
1290: \frac{kw-n}{2}~.
1291:  \nn
1292: \end{align}
1293: Denoting by $\Phi_{j;m,\bar{m}}$ the vertex operator with conformal weights
1294: $h= \frac{m^2-j(j+1)}{k}$, $\bar{h}= \frac{\bar{m}^2-j(j+1)}{k}$,
1295: %(whose ``Liouville part'' is  given as $\sim e^{-(j+1)\cQ \phi}$,
1296: %$\cQ=\sqrt{2/k}$, where $\phi$ is canonically normalized as
1297: %$\phi(z)\phi(0) \sim -\ln z$, $\Phi(\phi)
1298: %\sim - \frac{\cQ}{2}\phi,~
1299: %(\phi \rightarrow +\infty)$
1300: %).
1301: the exact reflection relation reads
1302: \begin{align}
1303: \Phi_{-(j+1);m,\bar{m}} = \cR(-(j+1),m,\bar{m}) \Phi_{j;m,\bar{m}}~, \label{qref rel}
1304: \end{align}
1305: The mini-superspace reflection amplitude $\cR_0(p,n)$ is
1306: then related to the exact one $\cR(j,m,\bar{m})$
1307: by taking the $k\,\rightarrow\,\infty$ limit as mentioned above
1308: (up to overall constant):
1309: \begin{align}
1310: \cR_0(p,n) = \lim_{k\rightarrow + \infty}\,
1311: \cR(j=-\frac{1}{2}+\frac{ip}{2}, m=\frac{n}{2}, \bar{m}=-\frac{n}{2})~.
1312: \end{align}
1313: \end{enumerate}
1314: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1315: 
1316: Although the mini-superspace approximation can only describe the momentum mode of the full spectrum, it is possible to study the winding mode by using the T-duality even if we restrict ourselves to the mini-superspace approximation. In the remaining part of this subsection, we will study the mini-superspace spectrum for the T-dualized trumpet geometry (i.e. $\bz_k$ orbifold of the vector coset $SL(2;\br)^{(V)}/U(1)$).
1317: T-dualized classical geometry is given by the trumpet geometry
1318: \begin{align}
1319: \dd s^2 = 2\left(k \dd\rho^2 + \frac{\dd\tilde{\theta}^2}{k\tanh^2\rho}\right) \ , \ \ e^{\Phi} = \frac{\mu}{\sinh\rho} \ . 
1320: \end{align}
1321: Note that we have a curvature and dilaton singularity at $\rho = 0$.
1322: For later purposes, let us discuss the minisuperspace analysis for the bulk spectrum. The minisuperspace spectrum is determined by the eigenfunctions of the string Laplacian
1323: \begin{align}
1324: \Delta &= - \frac{1}{e^{-2\Phi}\sqrt{\det G}} \partial_i e^{-2\Phi} \sqrt{\det{G}} G^{ij} \partial_j \cr
1325:  &= -\frac{2}{k}\left[\partial_\rho^2 + (\coth\rho+ \tanh\rho) \partial_\rho + k^2 \tanh^2\rho\partial_{\tilde{\theta}}^2 \right]
1326: \end{align}
1327: The (delta-function normalizable) eigenfunctions are given by
1328: \begin{align}
1329: \phi_{p,w}(\rho,\tilde{\theta}) &= C_1 e^{iw\theta}(\cosh\rho)^{-1-ip} F\left(\frac{1}{2}-\frac{kw}{2}+\frac{ip}{2},\frac{1}{2}+\frac{kw}{2}+ \frac{ip}{2};1+ip;\frac{1}{\cosh^2\rho}\right) \cr
1330:  &+ C_2 e^{iw\tilde{\theta}}(\cosh\rho)^{-1+ip} F\left(\frac{1}{2}-\frac{kw}{2}-\frac{ip}{2},\frac{1}{2}+\frac{kw}{2}- \frac{ip}{2};1-ip;\frac{1}{\cosh^2\rho}\right) \ .
1331: \end{align}
1332: It is not apriori clear which boundary condition one should impose because the trumpet geometry has a singularity at $\rho = 0$. Our natural guess would be to impose $\phi_{p,w}(\rho = 0,\tilde{\theta}) \equiv 0$. With our convenient normalization $C_1= 1$, this boundary condition amounts to 
1333: \begin{align}
1334: C_2 = \mathcal{R}_0(p,\omega) = \frac{\Gamma(ip)\Gamma(\frac{1}{2}-\frac{ip}{2}+\frac{kw}{2})\Gamma(\frac{1}{2}-\frac{ip}{2}-\frac{kw}{2})}{\Gamma(-ip)\Gamma(\frac{1}{2}+\frac{ip}{2}+\frac{kw}{2})\Gamma(\frac{1}{2}+\frac{ip}{2}-\frac{kw}{2})} \ , \label{ref dual}
1335: \end{align}
1336: which is consistent with the semiclassical limit of the exact reflection amplitude that is descended from the $SL(2;\br)$  WZNW model (or $\mathbb{H}_3^+$ model). Here we have used the formulae in the appendix A to evaluate the behavior of the hypergeometric function near the singularity.
1337: 
1338: 
1339: Before we close our study of the mini-superspace Euclidean two-dimensional black hole system, we introduce the so-called ``exact string background" for the bosonic two-dimensional black hole. As we have seen, for the bosonic string case, the spectrum shows $1/k$ corrections as
1340: \begin{align}
1341: h = \bar{h} = -\frac{j(j+1)}{k-2} + \frac{n^2}{4k} \ , \label{extst}
1342: \end{align}
1343: compared with the mini-superspace results
1344: \begin{align}
1345: h_0 = \bar{h}_0 =  -\frac{j(j+1)}{k} + \frac{n^2}{4k} \ .
1346: \end{align}
1347: To cure this small mismatch, \cite{Dijkgraaf:1992ba} introduced the following improved Laplacian
1348: \begin{align}
1349: \Delta'_0 = -\frac{1}{k-2}\left(\frac{\partial^2}{4\partial \rho^2} + \coth 2\rho \frac{\partial}{2\partial \rho} + (\coth^2\rho-\frac{2}{k})\frac{\partial^2}{\partial \theta^2}\right) \ , \label{imlap}
1350: \end{align}
1351: to reproduce the exact $1/k$ corrected spectrum \eqref{extst}. The corresponding metric is
1352: \begin{align}
1353: ds^2 = 2(k-2)\left(\dd\rho^2 + \frac{\dd\theta^2}{(\coth^2\rho -\frac{k}{2})} \right)
1354: \end{align}
1355: with the dilaton
1356: \begin{align}
1357: e^{2\Phi} = \mu \sinh2\rho\sqrt{\coth^2\rho-\frac{2}{k}} \ .
1358: \end{align}
1359: 
1360: In the literature, it has been shown that this background is a solution of the bosonic string equation of motion in a particular renormalization scheme \cite{Tseytlin:1992ri}. However, from the modern viewpoint, the reflection amplitude obtained from \eqref{imlap} is the same as the one from \eqref{Laplacian}, and does not capture the nonperturbative $1/k$ corrections that appear in the exact result \eqref{qref amp}. We will study the origin of the non-perturbative corrections to the mini-superspace reflection amplitude in section \ref{sec:3-4}. 
1361: 
1362: 
1363: 
1364: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1365: \subsection{Lorentzian spectrum}\label{sec:3-3}
1366: \subsubsection{classical string in two-dimensional black hole}\label{sec:3-3-1}
1367: We would like to study the classical string solution in the two-dimensional black hole geometry. For this purposes, we can directly solve the string equation of motion (and Virasoro constraint) on the classical background, or we can study the gauged WZNW model before integrating out the gauge constraint \cite{Bars:1994sv}.
1368: 
1369: If one takes the axial gauge $A=0$ in the classical gauged WZNW action \eqref{gWZW}, the solution can be constructed as follows.\footnote{Since we are studying the Lorentzian target space, we use the Lorentzian signature world-sheet.} The classical solution of the parent WZNW model is written as 
1370: \begin{align}
1371: g(\sigma_+,\sigma_-) =g_L(\sigma_+)g_R(\sigma_-)^{-1} \ , 
1372: \end{align}
1373: where we parametrize $g_L(\sigma_+)$, $g_R(\sigma_-)^{-1}$ $\in SL(2;\br)$ as
1374: \begin{align}
1375: g_L = \begin{pmatrix} a_L & u_L \\ -v_L  & b_L \end{pmatrix} \ ,  \ \ g_R^{-1} = \begin{pmatrix} b_R & -u_R \\ v_R  & a_R \end{pmatrix} \ , \\
1376: g =\begin{pmatrix} u_Lv_R+a_Lb_R & -a_Lu_R+u_La_R \\ b_Lv_R-v_Lb_R  & v_Lu_R+b_La_R \end{pmatrix}  \ ,
1377: \end{align}
1378: with the determinant constraint $u_Lv_L+a_Lb_L = u_Rv_R+a_Rb_R = 1$.
1379: 
1380: Now the current constraint $J_2 = \bar{J}_2=0$ reduces to
1381: \begin{align}
1382: \mathrm{Tr}(\sigma_3\partial_+ g_Lg_L^{-1})= v_L \partial_+ u_L + b_L \partial_+ a_L = -u_L\partial_+ v_L - a_L\partial_+ b_L = 0 \ , \label{curcon}
1383: \end{align}
1384: and the Virasoro constraint becomes\footnote{It is interesting to study general solutions of the coset CFT without imposing the Virasoro constraint. We will later come back to this point.}
1385: \begin{align}
1386: (a_L\partial_+ u_L - u_L \partial a_L)(b_L\partial_+v_L - v_L \partial_+ b_L) = 0 \ . \label{vircon}
1387: \end{align}
1388: The right moving part satisfies the similar equations.
1389: 
1390: Due to the determinant constraint, two equations in \eqref{curcon} are not independent, so we expect an appearance of one arbitrary function in the full solutions. Indeed, \eqref{curcon} and \eqref{vircon} suggests that either $\partial_+ u_L = \partial_+ a_L =0$, or $\partial_+v_L = \partial_+ b_L =0$ should be satisfied. Then the general solutions can be expressed as 
1391: \begin{align}
1392: g_L = \begin{pmatrix} a_L & u_L \\ -v_L(\sigma^+) &\frac{1}{a_L} - \frac{u_L}{a_L} v_L(\sigma^+)    \end{pmatrix} \ \text{or} \ \ \begin{pmatrix}  \frac{1}{b_L} - \frac{v_L}{b_L} u_L(\sigma^+) &u_L(\sigma^+) \\ -v_L &b_L   \end{pmatrix}  \cr
1393: g_R^{-1} =  \begin{pmatrix} \frac{1}{a_R} -\frac{u_R}{a_R} v_R(\sigma^-)  & -u_R \\ v_R(\sigma^-) & a_R\end{pmatrix} \ \text{or} \ \ \begin{pmatrix} b_R & -u_L(\sigma^-) \\ v_R  &\frac{1}{b_R} - \frac{v_R}{b_R} u_R(\sigma^-)  \end{pmatrix} \ .
1394: \end{align}
1395: Combining the left moving part and the right moving part, we totally obtain four possible solutions:
1396: \begin{align}
1397: g_A(\sigma_+,\sigma_-) &= \begin{pmatrix} \frac{1}{b}(1-{u}(\sigma^+){v}(\sigma^-)) & {u}(\sigma^+) \\ -{v}(\sigma^-) & b \end{pmatrix}  \cr
1398: g_B(\sigma_+,\sigma_-) &= \begin{pmatrix} a & \bar{u}(\sigma^-) \\ -\bar{v}(\sigma^+) & \frac{1}{a}(1-\bar{u}(\sigma^-)\bar{v}(\sigma^+)) \end{pmatrix} \cr
1399: g_C(\sigma_+,\sigma_-) &= \begin{pmatrix} {a}(\sigma^-) & c_1\\ \frac{1}{c_1}(-1+{a}(\sigma^{-}){b}(\sigma^+)) &{b}(\sigma^+) \end{pmatrix} \cr
1400: g_D(\sigma_+,\sigma_-) &= \begin{pmatrix} \bar{a}(\sigma^+) & \frac{1}{c_2}(1-\bar{a}(\sigma^+)\bar{b}(\sigma^-)) \\ -c_2 & \bar{b}(\sigma^-) \end{pmatrix} \ ,
1401: \end{align}
1402: where $u(\sigma^+), v(\sigma^-), \bar{u}(\sigma^{-}), \bar{v}(\sigma^{+}), a(\sigma^{-}), b(\sigma^+) , \bar{a}(\sigma^{+}), \bar{b}(\sigma^{-})$ are arbitrary real functions, and $(b,a,c_1,c_2)$ are real integration constants. We can read off the gauge invariant motion of strings from $u$ and $v$ components: 
1403: \begin{align}
1404: &A: &  u&=u(\sigma^+) \ , & v&= v(\sigma^-) \cr
1405: &B: &  u&=\bar{u}(\sigma^-) \ ,&  v&= \bar{v}(\sigma^+) \cr
1406: &C: &  u&=c_1 \ , & v&= \frac{1}{c_1}(1-a(\sigma^-)b(\sigma^+)) \cr
1407: &D: &  u&= \frac{1}{c_2}(1-\bar{a}(\sigma^+)\bar{b}(\sigma^-)) \ , & v&= c_2 \  . \label{sols}
1408: \end{align}
1409: It is interesting to note that the solution $A$ and $B$ are actually solutions of the string equation of motion  of any two-dimensional target space metric written in the conformal (Kruscal) coordinate: $\dd s^2 = f(u,v) \dd u\dd v$.
1410: 
1411: We still have a gauge degree of freedom associated with the conformal transformation: $\sigma^+ \to f(\sigma^+)$ and $\sigma^- \to \bar{f}(\sigma^-)$. By using this gauge symmetry, we can {\it locally} gauge away the arbitrary functions in  \eqref{sols} to make them reduce to the point particle (collapsed) string solution. For example, in the solution $A$ (similarly for $B$), we can expand
1412: \begin{align}
1413: u = u_0 + p_u(\tau + \sigma) + \sum_{n\neq 0} \alpha_n e^{-in(\tau + \sigma)} \ .
1414: \end{align}
1415: As is the case with the flat space, we can gauge away the oscillatory part, and due to the periodicity of $\sigma$ for closed strings, we have to set $p_u = 0$ unless the target space has a periodic directions. In the solution $C$ (similarly for $D$), we can locally set $ v = v(\tau)$ independent of $\sigma$ by using the conformal transformation. The resultant string motion is nothing but the geodesics for the massless point particle in the two-dimensional black hole. 
1416: 
1417: Thus the classical solution of the string equation collapses to a massless point particle in the two-dimensional black hole background, and it seems to be consistent with the vertex operator analysis, where the only tachyon fields are dynamical classically. However, if one allows a folded string solution, we can construct more general string solutions as we will review in the following section \ref{sec:3-3-2}. This can be done by patching different solutions of \eqref{sols} within the same world-sheet.
1418: 
1419: \subsubsection{folded strings}\label{sec:3-3-2}
1420: For an illustrative purpose, let us begin with the folded string solution  \cite{Bardeen:1975gx,Maldacena:2005hi} in two-dimensional flat space $(T,X)$. We can fix the world-sheet conformal invariance by choosing the gauge $\tau = T$.\footnote{This static gauge choice is not always possible especially in the curved space-time background.} The classical equation of motion is given by $\partial_+\partial_- X = 0$ with the Virasoro constraint\footnote{If one embeds the two-dimensional Minkowski space in higher dimensional space-time, the Virasoro constraint can be relaxed. For instance, if one introduces a contribution from the other zero modes than $(T,X)$ the resultant two-dimensional motion becomes massive rather than massless.}
1421: \begin{align}
1422: -2 + (\partial_+ X)^2 =  0  \ ,  \ \ -2 + (\partial_- X)^2 = 0 \ ,
1423: \end{align}
1424: which amounts to $\partial_+ X = \pm \sqrt{2}$. If we restrict ourselves to the solutions with continuous first derivative, they reduce to massless particles. To obtain the folded string solution, we can set $X= X_+(\sigma^+) + X_-(\sigma^-)$, where $X_+$ and $X_-$ are periodic functions with the same period and with derivatives $\pm \sqrt{2}$. In other words, we partition the world-sheet and assign different solutions that satisfy $ \partial_+ X = \pm \sqrt{2}$ and patch-work together so that the full solution is periodic in $\sigma$ direction. A simple solution with the static center of motion is given by 
1425: \begin{align}
1426: X = \sqrt{2}(|\sigma^+|_{\mathrm{per}} + |\sigma^-|_{\mathrm{per}}) \ , \label{minfold}
1427: \end{align}
1428: where the periodic absolute value function $|\sigma^+|_{\mathrm{per}}$ is given by $|\sigma^+|_{\mathrm{per}} = |\sigma^+|$ for $-\pi<\sigma^+ \le \pi$ and periodically extended outside this interval. More complicated solutions are possible, and one simple way to obtain them is to perform the target space Lorentz boost to the solution \eqref{minfold}. The resultant string solutions describes the folded string with the motion of the center of mass. Note that the Lorentz boost breaks our original gauge choice. 
1429: 
1430: Next. we will consider the linear dilaton background $\Phi = QX$. The equation of motion does not change, but the Virasoro constraint is modified as
1431: \begin{align}
1432: - 2 + (\partial_+ X)^2 - Q\partial_+^2 X = 0 \ ,  \ \ - 2 + (\partial_- X)^2 - Q\partial_-^2 X = 0 \ . 
1433: \end{align}
1434: The point is that due to the existence of the second derivatives in the Virasoro constraint, the solutions such as \eqref{minfold} are no more allowed. The most general solutions are given by
1435: \begin{align}
1436: X = X_0 -Q \log\left(\cosh\frac{\sqrt{2}\tau}{Q} + \cosh\frac{\sqrt{2}\sigma}{Q}\right) \ .
1437: \end{align}
1438: Except for the special limit where the string collapse to a point particle, the solution is not periodic in $\sigma$ direction. Thus only the long string stretched to the infinity is the allowed folded string solution. This agrees with the fact that there are no closed string states in the linear dilaton theory other than the massless tachyon. The inclusion of the Liouville potential does not alter the qualitative feature of the classical string solutions.
1439: 
1440: Let us now move on to the folded string solutions in the two-dimensional black hole. First, we partition the string world-sheet as in figure \ref{fig:partition}. We know that the solution should be locally given by \eqref{sols}. We fix the world-sheet conformal invariance by giving the boundary condition:
1441: \begin{align}
1442: &A: &  u&=u_0 +p^+\sigma^+_{\mathrm{per}}\ , & v&=v_0+p^-\sigma^-_{\mathrm{per}} \ ,\cr
1443: &B: &  u&=u_0 +p^+\sigma^-_{\mathrm{per}}\ , & v&=v_0+p^-\sigma^+_{\mathrm{per}}  \ , \label{first}
1444: \end{align}
1445: where $\sigma_{\mathrm{per}}^+ = \sigma^+ - m\pi$ for $\frac{\pi}{2}\le \sigma^+ <\frac{\pi}{2}$ and periodically identified outside this range.
1446: If we had considered a flat Minkowski space $\dd s^2 = \dd u\dd v$, the solution in the region $C$ and  $D$ would be independent of $\sigma$:
1447: \begin{align}
1448: &C_{\mathrm{flat}}: &  u&=u_0 + \frac{\pi p^{+}}{2}\ , & v&=v_0 + \frac{\pi p^-}{2}+ p^-(\sigma_{\mathrm{per}}^++\sigma_{\mathrm{per}}^-) \ \cr
1449: &D_{\mathrm{flat}}: &  u&=u_0 +\frac{\pi p^+}{2} +p^+(\sigma^-_{\mathrm{per}}+\sigma^-_{\mathrm{per}})\ , & v&=v_0+\frac{\pi p^-}{2}  \ .
1450: \end{align}
1451:  so that we have a fold as in \eqref{minfold}. In the two-dimensional black-hole case, the solution in $C$ and $D$ region is given by 
1452:  \begin{align}
1453:  C_{\mathrm{2DBH}}:  u&=u_0 + \frac{\pi p^{+}}{2}\ , \cr  v
1454:  &=\frac{1}{u_0+\frac{\pi p^+}{2}} \left(1-\frac{[1-(u_0+\frac{\pi p^+}{2})(v_0+p^-\sigma_{\mathrm{per}}^+)]\times{[1-(u_0+\frac{\pi p^+}{2})(v_0+p^-\sigma_{\mathrm{per}}^-)]}}{[1-(u_0+\frac{\pi p^+}{2})(v_0-\frac{\pi p^-}{2})]} \right) \ , \cr  
1455: D_{\mathrm{2DBH}}: u& =\frac{1}{v_0+\frac{\pi p^-}{2}} \left(1-\frac{[1-(v_0+\frac{\pi p^-}{2})(u_0+p^+\sigma_{\mathrm{per}}^+)]\times{[1-(v_0+\frac{\pi p^-}{2}(u_0+p^+\sigma_{\mathrm{per}}^-)]}}{[1-(v_0+\frac{\pi p^-}{2})(u_0-\frac{\pi p^+}{2})]} \right) \cr v&=v_0+\frac{\pi p^-}{2}  \ .
1456: \end{align}
1457: We could continue this analysis, but important point is that in the region $A'$ and $B'$ the solution turns out to be linear in $\sigma^+$ and $\sigma^-$ again just as in \eqref{first}. Thus the structure repeats itself and the simple recursion formula to derive the full solution exists (see \cite{Bars:1994sv}). Physically, the pulsating string falls into the black-hole as a massive particle. The folds move with the speed of light because $u$ or $v$ is constant.
1458: 
1459: \begin{figure}[htbp]
1460:    \begin{center}
1461:     \includegraphics[width=0.5\linewidth,keepaspectratio,clip]{partition.eps}
1462:     \end{center}
1463:     \caption{In order to obtain folded string solution, we partition the world-sheet and assign different solutions on each patch.}
1464:     \label{fig:partition}
1465: \end{figure}
1466: 
1467: 
1468: The folded string solution in the two-dimensional Liouville background was identified as the {\it open string} attached to the FZZT brane \cite{Fateev:2000ik,Teschner:2000md} in a certain limit \cite{Maldacena:2005hi}. The scattering amplitude computed from the CFT analysis completely matches with the matrix model computation in the non-singlet sector \cite{Fidkowski:2005ck,Kostov:2006dy}. In this sense, it makes sense to regard the non-singlet sector (or winding sector) in the matrix model corresponds to the folded string solution in the Lorentzian target space theory in the Liouville background.
1469: 
1470: Therefore, one might expect that the folded string solution in the two-dimensional black hole background should play an important role in constructing the dual matrix model for the two-dimensional black hole with the Lorentzian target space signature. At present, we do not have a conclusive argument for or against this direction, but we present some remarks here:
1471: \begin{itemize}
1472: 	\item It is important to note that the folded string solution in the Liouville background is in the open string sector and not in the closed string sector. This should be contrasted with the winding sector in the Euclidean two-dimensional black hole (and its hypothetical analytic continuation to the Lorentzian signature black hole).
1473: 	\item To obtain the long stretched string solution, the existence of the linear dilaton in the Virasoro constraint has been crucial. In our semiclassical analysis for the two-dimensional black hole, the existence of the nontrivial dilaton was neglected. This is the reason why we obtained a folded string solution that is asymptotically identical to the flat space solution \eqref{minfold}. Since in the linear dilaton case, such a solution was excluded, we might expect that only the long string solution would survive in the full quantization.\footnote{Just adding the dilaton term in the classical energy-momentum tensor is not consistent in the classical treatment because the one-loop correction to the equation of motion together with the classical contribution from the dilaton guarantee the one-loop holomorphic structure of the energy-momentum tensor.}
1474: 	\item In order to have a support for such long string solutions, we need a D-brane localized at asymptotic infinity. However, the two-dimensional Lorentzian  black hole does {\it not} admit such a D-brane solution from the classical DBI action analysis (see section \ref{sec:7-1}).
1475: \end{itemize}
1476: 
1477: We leave the role of the folded strings in the full quantization of the Lorentzian two-dimensional black hole system for future studies. In most of the following sections, we will concentrate on the mini-superspace approximation (point-particle approximation).
1478: 
1479: Before closing our discussion on the classical string solutions in the two-dimensional black hole, we would like to present the most general solutions of the  classical sigma model without imposing the Virasoro constraint for completeness \cite{deVega:1993pm}. First we introduce arbitrary three-component vectors lying on the hyperboloids
1480: \begin{align}
1481: \vec{A}\cdot\vec{A} &= - (A_0)^2 + (A_1)^2 + (A_2)^2 = 1 \cr
1482: \vec{B}\cdot\vec{B} &= - (B_0)^2 + (B_1)^2 + (B_2)^2 = 1 \ .
1483: \end{align}
1484: Then the most general solutions of the sigma model (in the Schwarzshild-like coordinate) are given by
1485: \begin{align}
1486: \cosh^2 \rho(\sigma,\tau) &= \frac{1}{2}\left(1+\vec{A}(\sigma^+)\cdot \vec{B}(\sigma^-)\right) \cr
1487: t(\sigma,\tau) &= \frac{1}{2}\int^{\sigma^+}_a\left(\frac{\epsilon_{abc}A'_aA_bB_c}{\vec{A}\cdot\vec{B}-1}\right)(x_+,b)dx_+ \cr
1488: &-\frac{1}{2}\int^{\sigma^-}_b\left(\frac{\epsilon_{abc}B'_aA_bB_c}{\vec{A}\cdot\vec{B}-1}\right)(\sigma_+,x_-)dx_- + c
1489: \end{align}
1490: The energy momentum tensor takes the form
1491: \begin{align}
1492: T_{++} = -\frac{A'(\sigma^+)^2}{8} \ ,  \ \ T_{--} = -\frac{B'(\sigma^-)^2}{8} \ .
1493: \end{align}
1494: If one imposes the Virasoro constraint, the solutions reduce to \eqref{sols}.
1495: 
1496: \subsubsection{spectrum from partition function?}\label{sec:3-3-3}
1497: The partition function for the Lorentzian two-dimensional black hole should be able to determine its spectrum in principle. However, in practice the subtleties associated with the  non-compactness of the target space and the Lorentzian signature make the partition function ill-defined and prevent us from reading the spectrum. 
1498: 
1499: At the classical level, the Lorentzian two-dimensional black hole is obtained by the Wick rotation $\theta \to i t$ in the cigar geometry. The vertex operator (corresponding to the massive character) should be Wick rotated as well
1500: \begin{align}
1501: \Delta(j,m=-\bar{m}=\frac{n}{2}) &= -\frac{j(j+1)}{k-2} + \frac{n^2}{4k} \cr
1502: \to \Delta(j,\omega) &= -\frac{j(j+1)}{k-2} - \frac{\omega^2}{4k}
1503: \end{align}
1504: based on the naive Wick rotation $n \to i \omega$. Since the time-direction $t$ is non-compact, $\omega$ should be continuous unlike the Euclidean two-dimensional black hole case, where $n$ is quantized. For the same reason, the Lorentzian black hole does not have a winding states along the $t$ direction.\footnote{Since the discrete states that descend from the $SL(2;\br)$ primaries have a winding quantum number in the Euclidean black hole, the corresponding states do not seem to exist in the Lorentzian black hole, but this is a controversial issue.} From the mini-superspace analysis we will review in \ref{sec:3-3-4}, the classical spectrum is obtained from this naive analytic continuation (with some care about the boundary conditions). 
1505: 
1506: We would like to make an attempt to read the spectrum from the Lorentzian partition function with some hindsight from the mini-superspace analysis.
1507: The proposed partition function (with a suitable analytic continuation) for the Lorentzian $SL(2;\br)/U(1)$ coset is 
1508: \begin{align}
1509: Z_{SL(2;\br)/U(1)} =  \int_{\br^2} \frac{\dd v^2}{\tau_2}\frac{e^{\frac{v_1^2}{\tau_2}-\frac{\pi k}{\tau_2}|v|^2}  }{\sqrt{\tau_2}|\theta_1(\tau,iv)|^2 } \sqrt{\tau_2}|\eta(\tau)|^2  \ . \label{ppl}
1510: \end{align}
1511: Here we study the bosonic case first for simplicity. Since we are integrating over the whole complex plane spanned by $v$, this partition function is formally same as that for the Euclidean axial coset \eqref{ppo}. The conclusion that the spectrum of the Euclidean coset and the Lorentzian coset, nevertheless, has the same spectrum seems too quick given the fact that the mini-superspace approximation gives a totally different answer. 
1512: 
1513: In the Euclidean case, the divergence near $iv = u = n+ \omega \tau$ gives a bulk contribution of the noncompact boson (with a linear dilaton) coupled to a compact boson with radius $R^2 = k$, where the summation over $n$ and $\omega$ shows the existence of the compact direction. In the Lorentzian case, we propose that the the torus modulus $\tau$ should be Lorentzian, namely $\tau$ and $\bar{\tau}$ and real and mutually independent. 
1514: 
1515: 
1516: On the Lorentzian torus, the divergence of the partition function \eqref{ppl} only appears at $v=0$, leading to the contribution of the noncompact boson (with a linear dilaton) coupled to a non-compact boson:\footnote{On the Lorentzian torus, the Euclidean coset partition function still has an infinitely many origin of divergence at $ u = n+\omega \tau$, which gives a compact boson.}
1517: \begin{align}
1518: Z \sim \log \epsilon Z_{free}^2 + \text{finite terms}\ .
1519: \end{align}
1520: The leading order partition function seems to agree with the mini-superspace analysis.\footnote{One should note that the partition function on the Lorentzian torus needs a careful $i\epsilon$ prescription. This is an interesting but subtle subject, and we will not delve into the details here.}
1521: 
1522: More precise discussions are needed to determine the finite part of the partition function. The situation is more involved than in the Euclidean case, and so far no conclusive agreements are available.
1523: One interesting related question is how the target-space supersymmetry is broken in the (world-sheet supersymmetric) Lorentzian two-dimensional black hole background. If we restrict ourselves to the leading order partition function, we can certainly define a GSO projection, and the partition function (at the leading order) vanishes because asymptotically the theory is essentially free and the spectrum coincides with the supersymmetric linear dilaton theory, and the target-space supersymmetry operator can be constructed. Therefore, the target-space supersymmetry should be broken near the horizon, where the curvature effects will be present.
1524: 
1525: To study this problem, let us consider the type II GSO projected partition function for the two-dimensional Lorentzian black hole (coupled to free transverse sectors that is represented by free CFTs for simplicity). Our original partition function for the Lorentzian $SL(2;\br)/U(1)$ Kazama-Suzuki coset is the diagonal modular invariant one:
1526: \begin{align}
1527: Z^{(NS)}(\tau) = \int_{\br^2} \frac{\dd v^2}{\tau_2} \frac{|\theta_3(\tau,iv)|^2}{\sqrt{\tau_2}|\theta_1(\tau,iv)|^2 } \sqrt{\tau_2}|\eta(\tau)|^2 e^{-\frac{\pi k}{\tau_2}|v|^2} \ 
1528: \end{align}
1529: for NS-NS sectors (other sectors can be obtain by replacing $\theta_3(\tau,u)$ with $\theta_a(\tau,u)$ $(a=0,1,2)$ from the spectral flow). A natural candidate for the type II partition function is 
1530: \begin{align}
1531: Z(\tau) 
1532: = \int_{\br^2} \frac{\dd v^2}{\tau_2} \frac{|\theta_3(\tau,iv)\theta_3^3 - \theta_2(\tau,iv)\theta_2^3 \pm \theta_1(\tau,iv)\theta_1^3 - \theta_0(\tau,iv)\theta_0^3)|^2}{\sqrt{\tau_2}|\theta_1(\tau,iv)|^2 } \sqrt{\tau_2}|\eta(\tau)|^{-14} e^{-\frac{\pi k}{\tau_2}|v|^2} \ , \label{typii}
1533: \end{align}
1534: where $\theta_a^3 = \theta_a(\tau,0)^3$ have been introduced from the free CFT contributions. The expression is almost supersymmetric: for example, if we first take the leading diverging part at $v=0$, then the fermionic oscillator part gives zero after summing over the spin structure due to the abstruse identity of Jacobi.
1535: 
1536: A remaining uncancelled part, which describes the breaking of the target-space supersymmetry, is of particular interest.
1537: By using the Riemann quartic identity \eqref{rqi}, we can rewrite \eqref{typii} as
1538: \begin{align}
1539: Z(\tau) = \int_{\br^2} \frac{\dd v^2}{\tau_2} \frac{|\theta_1(\tau,i\frac{v}{2})|^8}{\sqrt{\tau_2}|\theta_1(\tau,iv)|^2 } \sqrt{\tau_2}|\eta(\tau)|^{-14} e^{-\frac{\pi k}{\tau_2}|v|^2} \ . \label{typeiit}
1540: \end{align}
1541: Now one can see that the leading order divergence near the origin $(v=0)$ is indeed removed, which suggests that the bulk part of the spectrum is supersymmetric.\footnote{Unfortunately, for complex $\tau$, the partition function still shows a volume divergence at $iv = n + \omega \tau$ with a pair of {\it odd} integers $(n,\omega)$, where the GSO projection acts {\it oppositely}. We do not fully understand the origin of this failure of bulk cancellation. Since these divergences seem unphysical in the Lorentzian partition function if we stick to the Lorentzian torus, we do not see their physical relevance.}
1542: 
1543: It is still difficult to evaluate \eqref{typeiit} to uncover the non-supersymmetric spectrum of the two-dimensional Lorentzian black hole partially because the formal $q$ expansion of \eqref{typeiit} gives a divergent series. Let us suppose that the major part of the $v$ integral comes near the origin $v=0$ since in the large $k$ limit, the Gaussian factor would provide a strong convergence factor for the integral (with no good justification because of the oscillatory nature of the $\theta$ functions on the Lorentzian torus). The subsequent integration over $v$ would lead to
1544: \begin{align}
1545: Z(\tau) \sim |\eta(\tau)|^4 \ .
1546: \end{align}
1547: The partition function looks like a free 0-dimensional bosonic string (probably localized near the horizon). The support of the (non-supersymmetric) bosonic degrees of freedom should be localized because we do not have a diverging volume factor.
1548: 
1549: It seems likely that the supersymmetry is broken only locally in the vicinity of the horizon. One naively guess that this should correspond to the Lorentzian version of the winding condensation near the tip of the cigar in the Euclidean two-dimensional black hole as we will see in section \ref{sec:3-4}. A further study of this subject is of great interest and the more precise definition of the (almost) supersymmetric partition function \label{typeiit} and its evaluation is highly desirable.
1550: 
1551: \subsubsection{minisuperspace approximation}\label{sec:3-3-4}
1552: Since it is difficult to read the spectrum of the Lorentzian two-dimensional black hole directly from the partition function, it is very important to study the classical spectrum based on the mini-superspace approximation.
1553: The Wick rotation of the mini-superspace eigenfunctions in the
1554: Euclidean cigar geometry \eqref{ef} is not so trivial. Fortuitously,
1555: the Lorentzian eigenfunctions are already classified thoroughly in
1556: \cite{Dijkgraaf:1992ba}. The complete basis for waves outside the black hole
1557: horizon are spanned by the following four types of
1558: eigenfunctions\footnote{Here we adopt slightly different
1559: normalization  from \cite{Dijkgraaf:1992ba}.} of the Lorentzian
1560: Klein-Gordon operator. For those with the eigenvalue
1561: $\frac{p^2}{4k}-\frac{\om^2}{4k}+ \frac{1}{4k}$ of the Klein-Gordon
1562: operator, the four eigenfunctions are
1563: \begin{align}
1564: U^p_{\om}(\rho,t) &= -
1565: \frac{\Gamma^2(\nu_+)}{\Gamma(1-i\om)\Gamma(-ip)} e^{-i\om t} (\sinh
1566: \rho)^{-i\om} F(\nu_+,\nu^*_-;1-i\om;-\sinh^2\rho) \nn &\sim
1567: e^{-i\om t - i\om \ln \rho} \qquad \mbox{as} \qquad
1568: \rho\,\rightarrow\,0~,
1569: \label{U} \\
1570: V^p_{\om}(\rho,t) &=  -
1571: \frac{\Gamma^2(\nu^*_+)}{\Gamma(1+i\om)\Gamma(ip)} e^{-i\om t}
1572: (\sinh \rho)^{i\om} F(\nu^*_+,\nu_-;1+i\om;-\sinh^2\rho) \nn &\sim
1573: e^{-i\om t + i\om \ln \rho} \qquad \mbox{as} \qquad
1574: \rho\,\rightarrow\,0~,
1575: \label{V} \\
1576: L^p_{\om} (\rho,t ) &= e^{-i\om t} (\sinh \rho)^{-1-ip} F(\nu^*_+,
1577: \nu^*_-;1+ip; -\frac{1}{\sinh^2 \rho}) \nn &\sim e^{-\rho}
1578: e^{-ip\rho -i\om t} \qquad \mbox{as} \qquad
1579: \rho\,\rightarrow\,\infty~,
1580: \label{L} \\
1581: R^p_{\om} (\rho,t )
1582: %(\equiv L^{-p}_{\om}(\rho,t))
1583: &= e^{-i\om t} (\sinh \rho)^{-1+ip} F(\nu_+, \nu_-;1-ip;
1584: -\frac{1}{\sinh^2 \rho}) \nn &\sim e^{-\rho} e^{+ip\rho -i\om t}
1585: \qquad \mbox{as} \qquad \rho\,\rightarrow\,\infty \label{R}
1586: \end{align}
1587: with the notations
1588: \begin{align}
1589: \nu_{\pm} = \frac{1}{2} - i\left(\frac{p}{2}\pm
1590: \frac{\om}{2}\right)~. \nonumber
1591: \end{align}
1592: These eigenfunctions are defined by the following analytic
1593: continuations of the mini-superspace Euclidean eigenfunctions:
1594: \begin{align}
1595:  U^p_{\om}(\rho,t)& =
1596: \left\{
1597: \begin{array}{ll}
1598: \phi^p_{n=+i\om}(\rho,\theta = +it)~  ~~ (\om>0,~n<0)  \\
1599: \phi^p_{n=-i\om}(\rho,\theta = -it)~  ~~ (\om<0,~n>0)
1600: \end{array}
1601: \right. \nn
1602:  V^p_{\om}(\rho,t&)=
1603: \left\{
1604: \begin{array}{ll}
1605: \phi^{-p}_{n=-i\om}(\rho,\theta = - it)~  ~~ (\om>0,~n<0)  \\
1606: \phi^{-p}_{n=+i\om}(\rho,\theta = +it)~  ~~ (\om<0,~n>0)
1607: \end{array}
1608: \right. \nn  L^p_{\om}(\rho,t)&= \phi^p_{L,n=i\om} (\rho,
1609: \theta=+it) \nn  R^p_{\om}(\rho,t)&= \phi^p_{R,n=i\om} (\rho,
1610: \theta=+it) ~, \label{ac UVLR}
1611: \end{align}
1612: %
1613: where the $n<0$ and $n>0$ ranges are mapped to $\om>0$ and $\om<0$,
1614: respectively.
1615: 
1616: As discussed in \cite{Dijkgraaf:1992ba}, only two out of the four eigenfunctions
1617: are linearly independent. In particular,
1618: %
1619: \begin{align}
1620: V_\omega^p(\rho, t) = U^{p*}_\omega(\rho, -t) \qquad \mbox{and}
1621: \qquad R_\omega^p(\rho, t) = L^{p*}_\omega (\rho, -t). \nonumber
1622: \end{align}
1623: %
1624: The reason why we introduce the above four eigenfunctions is because
1625: they encode four possible boundary conditions (We here assume $p>0$)
1626: in the Lorentzian black hole background. Recall that, for the region
1627: outside the horizon of the eternal black hole, the boundaries
1628: consist of four segments: `future (past) horizon' $t=+\infty, \,
1629: \rho=0$ ($t=-\infty, \, \rho=0$) by ${\cal H}^+$ (${\cal H}^-$), and
1630: the `future (past) infinity' $t=+\infty, \, \rho=+\infty$
1631: ($t=-\infty, \, \rho=+\infty$) by ${\cal I}^+$ (${\cal I}^-$). The
1632: four eigenfunctions $U, V, L, R$ are the ones obeying boundary
1633: conditions:
1634: \begin{align}
1635:  U^p_{\om} &= 0 ~~ \mbox{at}~ {\cal H}^-~,~~~ V^p_{\om} = 0 ~~
1636: \mbox{at}~~ {\cal H}^+~, \cr ~~~ L^p_{\om} &= 0~ (R^p_{\om} = 0)~~ 
1637: \mbox{at}~~ {\cal I}^+~, ~~~ R^p_{\om} = 0 ~(L^p_{\om} = 0) ~~
1638: \mbox{at}~~ {\cal I}^- %\nonumber
1639: %\label{bd ef}
1640: \end{align}
1641: for $\om>0$ ($\om <0$). See Figure \ref{wave}.
1642: 
1643: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1644: %
1645: \begin{figure}[htbp]
1646:     \begin{center}
1647:  %   \includegraphics[width=0.6\linewidth,keepaspectratio,clip]
1648:    \includegraphics[width=13cm,height=10cm]
1649:       {wave.eps}
1650:     \end{center}
1651:     \caption{The boundary conditions of the Lorentzian
1652:        eigenfunctions ($\om>0$ sector). For $\om<0$, the
1653:         figures for $L$ and $R$ should be interchanged.}
1654:     \label{wave}
1655: \end{figure}
1656: 
1657: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1658: 
1659: By Wick rotating the mini-superspace reflection relations \eqref{cref
1660: rel}, we obtain linear relations among the Lorentzian
1661: eigenfunctions:
1662: \begin{align}
1663:  U^p_{\om} = L^{p}_{\om} + \cR_0(p,\om) R^p_{\om} \qquad
1664: \mbox{and} \qquad V^p_{\om} = R^{p}_{\om} + \cR^*_0(p,\om)
1665: L^p_{\om}~. \label{decomp ef 2}
1666: \end{align}
1667: Equivalently,
1668: \begin{align}
1669:  L^p_{\om} &= \frac{1}{1-|\cR_0(p,\om)|^2} \left\{ U^p_{\om} -
1670: \cR_0(p,\om) V^p_{\om} \right\} \cr \quad \mbox{and} \quad R^p_{\om} &=
1671: \frac{1}{1-|\cR_0(p,\om)|^2} \left\{ V^p_{\om} - \cR^*_0(p,\om)
1672: U^p_{\om} \right\}. \label{decomp ef 2-2}
1673: \end{align}
1674: Here, the mini-superspace reflection amplitude $\cR_0(p,\om)$ in
1675: Lorentzian theory is given by
1676: \begin{align}
1677: \cR_0(p,\om) = \frac{\Gamma(+ip)\Gamma^2(\nu_+)}
1678: {\Gamma(-ip)\Gamma^2(\nu^*_-)}
1679: %\nn
1680: % &=& \frac{\Gamma(ip)\Gamma(\nu_+)\Gamma(\nu_-)}
1681: %{\Gamma(-ip)\Gamma(\nu^*_+)\Gamma(\nu^*_-)}
1682: %\times \frac{\cosh \pi \left(\frac{p-|\om|}{2}\right)}
1683: %{\cosh \pi \left(\frac{p+|\om|}{2}\right)} \nn
1684:  \equiv - \frac{B(\nu_+,\nu_-)}{B(\nu_+^*,\nu_-^*)} \cdot
1685: \frac{\cosh \pi \left(\frac{p-\om}{2}\right)}
1686: {\cosh \pi \left(\frac{p+\om}{2}\right)}~.
1687: \label{cref amp 2}
1688: \end{align}
1689: Notice that, in sharp contrast to the Euclidean black hole, the
1690: reflection amplitude is less than unity due to the second factor:
1691: %
1692: \begin{align} |\cR_0(p,\om)|^2 = {\cosh^2 \pi \left({p-\om \over 2}\right)
1693: \over \cosh^2 \pi \left({p+\om \over 2} \right)} \leq 1.
1694: \label{inequality} \end{align}
1695: %
1696: The inequality is saturated at $p=\omega = 0$. The inequality
1697: \eqref{inequality} shall play a prominent role for understanding string
1698: dynamics in the Lorentzian black hole background. The
1699: mini-superspace reflection relations for $U^p_{\om}$, $V^p_{\om}$
1700: are also expressible in a form similar to the Euclidean ones.
1701: Recalling that $\cR_0(-p, \om) \cR_0(+p, \om) = 1$,
1702: \begin{align}
1703: U^{-p}_{\om}(\rho,t)= \cR_0(-p,\om) U^{p}_{\om}(\rho,t) \qquad
1704: \mbox{and} \qquad V^{-p}_{\om}(\rho,t)= \cR^*_0(-p,\om)
1705: V^{p}_{\om}(\rho,t)~, \label{cref rel UV}
1706: \end{align}
1707: while $L^p_{\om}$ and $R^p_{\om}$ are simply related by reflection:
1708: \begin{align}
1709:  L^{-p}_{\om} (\rho,t) = R^{+p}_{\om} (\rho,t)~. \label{cref rel
1710: LR}
1711: \end{align}
1712: Moreover, $U^p_{\om}$ and $V^p_{\om}$ are linearly independent
1713: except for the special kinematic regime, $\om=0$. Notice also, in
1714: the relation \eqref{inequality}, the reflection amplitude involves the
1715: mini-superspace contribution only, not the full-fledged stringy one.
1716: 
1717: %%%%%%%%%%
1718: Before proceeding further, we shall here collect explicitly
1719: relations among inner products of Lorentzian primary fields, where
1720: the inner product is defined with respect to the Lorentzian measure
1721: $\dd v_L = k\sinh 2\rho \dd \rho \dd t$. Taking quantum
1722: numbers $p$, $\om$ fixed and dropping off delta function factors
1723: $2\pi\delta(p-p')$, $2\pi\delta(\om-\om')$ for notational
1724: simplicity, we have
1725: \begin{align}
1726:  & (U^p_{\om}, U^p_{\om}) = (V^p_{\om}, V^p_{\om})
1727:    = N_0(p,\om)~, \qquad N_0(p,\om)
1728: \equiv \frac{1+|\cR_0(p,\om)|^2}{2}
1729: \nn
1730:  & (U^p_{\om},V^p_{\om}) = \cR_0^*(p,\om)~,
1731: \nn
1732:  & (L_{\om}^p,L_{\om}^p)= (R^p_{\om}, R^p_{\om})= \frac{1}{2}~,
1733:  \qquad \quad (L^p_{\om}, R^p_{\om})= 0 ~, \nn
1734:  & (U^p_{\om},L^p_{\om}) = (V^p_{\om}, R^p_{\om}) =\frac{1}{2}~,
1735:  \qquad \quad (R^p_{\om},U^p_{\om}) = (V^p_{\om}, L^p_{\om})
1736:    =\frac{\cR_0(p,\om)}{2}~.
1737: \label{inner product UVLR}
1738: \end{align}
1739: The inner products involving $L^p_\om$ and $R^p_\om$ are readily
1740: evaluated since dominant contributions are supported in the
1741: asymptotic region $\rho \gg 0$, yielding the volume factor
1742: $2\pi\delta(0)$. The remaining inner products ca be extracted
1743: from the linear relations \eqref{decomp ef 2}, \eqref{decomp ef 2-2}.\footnote
1744: {We checked these inner products numerically using
1745: MATHEMATICA.} We also fixed the overall normalization factors from
1746: consistency with the Euclidean inner product \eqref{inner product}
1747: under the $\om\,\rightarrow\, 0$ limit. Notice also that
1748: \begin{align}
1749:  N_0(-p,\om) = \left|\cR_0(-p,\om)\right|^2 \, N_0(+p,\om)~,
1750: \end{align}
1751: as is consistent with the mini-superspace reflection relation
1752: \eqref{cref rel UV}.
1753: 
1754: 
1755: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1756: It is easy to construct the exact string vertex operators or primary
1757: states corresponding to the mini-superspace eigenfunctions $U$, $V$,
1758: $L$, $R$.
1759: %%%
1760: To be specific, we shall consider primarily the fermionic $SL_k (2,
1761: \br)/U(1)$ supercoset conformal field theory.\footnote
1762:     {For the bosonic $SL(2;\br)_{\kappa}/U(1)$ coset conformal field theory,
1763:   we instead have
1764: $h= \bar{h}= \frac{p^2}{4(\kappa-2)}-\frac{\om^2}{4\kappa}
1765:  +\frac{1}{4(\kappa-2)}$, and
1766: $\cR(p,\om)\equiv
1767: \cR_0(p,\om) \frac{\Gamma\left(1+\frac{ip}{\kappa-2}\right)}
1768: {\Gamma\left(1-\frac{ip}{\kappa-2}\right)}$.
1769: }
1770: %%%
1771: The primary states $\ket{U^p_{\om}}$, $\ket{V^p_{\om}}$ are the ones
1772: of conformal weights $h= \bar{h}= \frac{p^2}{4k}-\frac{\om^2}{4k}
1773: +\frac{1}{4k}$ and obey the exact reflection relations
1774: \begin{align}
1775:  \ket{U^{-p}_{\om}}= \cR(-p,\om) \ket{U^{p}_{\om}}~, ~~~
1776: \ket{V^{-p}_{\om}}= \cR^*(-p,\om) \ket{V^{p}_{\om}}~, \label{qref rel
1777: UV}
1778: \end{align}
1779: and the exact reflection amplitude is given by
1780: %
1781: \begin{align} \cR(p,\om)\equiv \cR_0(p,\om)
1782: \frac{\Gamma\Big(1+\frac{ip}{k}\Big)}
1783: {\Gamma\Big(1-\frac{ip}{k}\Big)}. \label{exactra} \end{align}
1784: %
1785: Notice that the string world-sheet effect entering through the
1786: $1/k$-correction is a pure phase. Thus, the exact reflection
1787: probability $\vert {\cal R} (p, \om) \vert^2$ remains unmodified
1788: from the mini-superspace approximation result $\vert {\cal R}_0(p,
1789: \om) \vert^2$ given in \eqref{inequality}. We shall normalize the
1790: primary states $\ket{U^p_{\om}}$, $\ket{V^p_{\om}}$ ($p>0$) as
1791: \begin{align}
1792: & \bra{U^p_{\om}} U^{p'}_{\om'}\rangle = \bra{V^p_{\om}}
1793: V^{p'}_{\om'}\rangle = N(p,\om) \, 2\pi\delta(p-p')
1794: 2\pi\delta(\om-\om') ~, \nonumber \\
1795: & \bra{V^p_{\om}} U^{p'}_{\om'} \rangle = \cR^* (p,\om) \,
1796: 2\pi\delta(p-p') 2\pi\delta(\om-\om') ~, \label{norm UV}
1797: \end{align}
1798: where the new normalization factor $N(p,\om)$ is simply defined by
1799: replacing $\cR_0$ with $\cR$ in $N_0(p,\om)$. The primary states
1800: $\ket{L^p_{\om}}$, $\ket{R^p_{\om}}$ are also definable by using the
1801: linear relations \eqref{decomp ef 2} or \eqref{decomp ef 2-2} but now
1802: with $\cR_0$ replaced by $\cR$. Notice that $\ket{U^p_{\om}}$,
1803: $\ket{V^p_{\om}}$ are the ones analytically continuable to the
1804: Euclidean primary states $\ket{\phi^{\pm p}_n}$, so often referred
1805: as the `Hartle-Hawking vacua'. On the other hand, the states
1806: $\ket{L^p_{\om}}$, $\ket{R^p_{\om}}$ does not have Euclidean
1807: counterparts. Recall that, over the Euclidean black hole background,
1808: $\phi^p_{L,n}$, $\phi^p_{R,n}$ behave badly in the vicinity of $\rho
1809: = 0$ and hence ill-defined.
1810: 
1811: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1812: 
1813: We also find it useful to introduce the dual basis
1814: $\widehat{\bra{U^p_{\om}}}$, $\widehat{\bra{V^p_{\om}}}$ ($p,p'>0$)
1815: with inner products
1816: \begin{align}
1817:  & \widehat{\bra{U^p_{\om}}} U^{p'}_{\om'} \rangle
1818:    = \widehat{\bra{V^p_{\om}}} V^{p'}_{\om'} \rangle
1819:    = 2\pi\delta(p-p')2\pi\delta(\om-\om')~, \qquad
1820:  \widehat{\bra{U^p_{\om}}} V^{p'}_{\om'} \rangle
1821:    = \widehat{\bra{V^p_{\om}}} U^{p'}_{\om'} \rangle
1822:    = 0~.
1823: \label{hat U V}
1824: \end{align}
1825: Explicitly, they are given by
1826: \begin{align}
1827:  \widehat{\bra{U^p_{\om}}} &= \frac{2}{1-\left|\cR(p, \om)
1828: \right|^2} \left\{ \bra{L^p_{\om}} - \cR^* (p, \om) \bra{R^p_{\om}}
1829: \right\}~, \cr \qquad \widehat{\bra{V^p_{\om}}} &=
1830: \frac{2}{1-\left|\cR(p, \om) \right|^2} \left\{ \bra{R^p_{\om}} -
1831: \cR(p, \om) \bra{L^p_{\om}} \right\}~. \label{def hat U V}
1832: \end{align}
1833: %
1834: As such, these dual basis obey the following exact reflection
1835: relations:
1836: \begin{align}
1837:  \widehat{\bra{U^{-p}_{\om}}} = \cR(p,\om)
1838:  \widehat{\bra{U^{p}_{\om}}} \qquad \mbox{and} \qquad
1839: \widehat{\bra{V^{-p}_{\om}}} = \cR(p,\om)^*
1840:  \widehat{\bra{V^{p}_{\om}}}~.
1841: \label{ref hat U V}
1842: \end{align}
1843: 
1844: A remark is in order. The dual basis $\widehat{\bra{U^{p}_{\om}}}$,
1845: $\widehat{\bra{V^{p}_{\om}}}$ are {\em not\/} Wick rotatable to the
1846: Euclidean dual basis $\bra{\phi^{+p}_n}$, $\bra{\phi^{-p}_n}$, since
1847: $|\cR(p,\om)|=1$ for $\om \in i\br$. The correct procedure would be
1848: that we first define Wick rotations for the `ket' states, and then
1849: define their dual states within the Lorentzian Hilbert space.
1850: Nevertheless, one-point correlators in the Lorentzian theory, from
1851: which a set of physical observables can be computed, ought to be
1852: always analytically continuable to the one-point correlators in the
1853: Euclidean theory. Roughly speaking, ambiguities inherent to the Wick
1854: rotation of dual states drop out upon taking inner product.
1855: %%%%
1856: %
1857: %Having obtained the Lorentzian primary states, we shall now
1858: %construct several interesting class of boundary states for a
1859: %D0-brane propagating in the black hole background. We have seen that
1860: %the D0-brane propagates along the trajectory \eqref{trajectory D0}.
1861: %The two-dimensional black hole is eternal, so, in addition to the
1862: %past and the future asymptotic infinities, the causal propagation
1863: %region has the past horizon ${\cal H}^-$ surrounding the white hole
1864: %singularity and the future horizon ${\cal H}^+$ surrounding the
1865: %black hole singularity. As such, by taking variety of possible
1866: %boundary conditions, we can construct interesting class of boundary
1867: %states.
1868: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1869: \subsection{Duality and winding tachyon condensation}\label{sec:3-4}
1870: One of the salient features of the (Euclidean) $SL(2;\br)/U(1)$ coset model is the so-called Fateev-Zamolodchikov-Zamolodchikov (FZZ) duality \cite{FZZ}. Mathematically speaking, this duality has enabled us to compute exact two-, and three-point functions of the $SL(2;\br)/U(1)$ coset model and revealed their pole structures. Physically speaking, on the other hand, it has established a duality between the winding tachyon condensation and the singularity resolution of the geometry, and uncovered, from an exact CFT perspective, the importance of the winding tachyon condensation near the classical singularities.
1871: 
1872: Let us formulate the FZZ duality in the $\mathcal{N}=2$ supersymmetric case. The FZZ duality states:
1873: 
1874: {\bf FZZ duality} ($\mathcal{N}=2$) supersymmetric $SL(2;\br)/U(1)$ coset model with level $k$ is equivalent (up to chirality) to the $\mathcal{N}=2$ Liouville field theory (see e.g. \cite{Nakayama:2004vk} for reference): 
1875: \begin{align}
1876: L = \int \dd^4\theta \Phi^\dagger \Phi + \int \dd^2\theta W(\Phi) + h.c. \cr
1877: W(\Phi) = \mu e^{\frac{1}{Q}\Phi}  \ ,
1878: \end{align}
1879: where $Q^2 = \frac{2}{k}$.
1880: 
1881: 
1882: The appearance of the chirality flip suggests the T-dual nature of the duality. Indeed, the FZZ duality can be proved in a more general context of the mirror symmetry. Physically, the appearance of the $\mathcal{N}=2$ Liouville potential in the T-dualized set up can be anticipated as follows. As we studied in section \ref{sec:3-1}, the T-dual of the $SL(2;\br)^{(A)}/U(1)$ axial coset model, whose classical geometry is the cigar, is classically described by the trumpet geometry. However, the trumpet geometry has a singularity coming from the fixed point of the (T-dualized) $U(1)$ angular direction. To avoid the existence of a naked singularity of the space-time, the (T-dualized) winding tachyon will condensate. From the world-sheet viewpoint, the (winding) tachyon condensation is nothing but the $\mathcal{N}=2$ Liouville superpotential.\footnote{To avoid a possible confusion, we note that the original winding tachyon condensation becomes non-winding tachyon with $\tilde{\theta}$ momentum after taking the T-duality.}
1883: 
1884: The operator correspondence of the FZZ duality is almost clear. In the asymptotic region, one can write the vertex operators of primary states in the $SL(2;\br)^{(A)}/U(1)$ coset model by using those of the linear dilaton times $U(1)$ angular direction. We then perform the T-duality, to write them down as asymptotic vertex operators of primary states in the $\mathcal{N}=2$ Liouville theory. The descendant structure is completely fixed by the $\mathcal{N}=2$ superconformal algebra.
1885: 
1886: There are several ``proofs" of the FZZ duality available in the literature. In the original work, FZZ has given a direct computation of the two- and three-point functions of the both models (including winding violating correlation functions) and has shown the equivalence between the two models when the computation based on the screening operator is available. In \cite{Hori:2001ax}, the duality has been established rigorously at the level of the topological field theory from the viewpoint of the mirror symmetry (T-duality of the linear sigma model that flows to $SL(2;\br)/U(1)$ coset in the infrared). As is the case with the usual mirror symmetry, it is natural to expect that the full conformal field theory is dual with each other, and indeed there is much supporting evidence for that. In another interesting derivation of the FZZ duality \cite{Tong:2003ik}, the domain wall dynamics of  a certain $2+1$ dimensional gauge theory has been studied, resulting in two complementary descriptions --- $SL(2;\br)/U(1)$ coset model on one hand and $\mathcal{N}=2$ Liouville theory on the other hand.
1887: 
1888: We will not review the derivation of the FZZ duality (see any of the references above, or consult \cite{Nakayama:2004vk} for a brief summary of the related discussions). Instead, we will see some physical consequences of the duality in the remaining part of this section. Let us begin with the comparison between the classical two-point function of the $SL(2;\br)/U(1)$ coset CFT from the minisuperspace approximation and the exact one. The mini-superspace result (see \eqref{cref amp} and \eqref{ref dual}) is 
1889: \begin{align}
1890: \cR_0(j,m,\bar{m}) =
1891: \frac{\Gamma(2j+1)\Gamma(-j+m)\Gamma(-j-\bar{m})}{\Gamma(-2j-1)
1892: \Gamma(j+1+m)\Gamma(j+1-\bar{m})}, \end{align}
1893: %
1894: where
1895: %
1896: \begin{align}  \qquad m=\frac{kw+n}{2}~, \qquad \bar{m} =
1897: \frac{kw-n}{2}~ ,
1898:  \nn
1899: \end{align}
1900: while the exact result is 
1901: \begin{align}
1902: \cR(j,m,\bar{m}) &= \nu(k)^{-2j-1}\,
1903: \frac{\Gamma(1+\frac{2j+1}{k})}{\Gamma(1-\frac{2j+1}{k})}
1904: \frac{\Gamma(2j+1)\Gamma(-j+m)\Gamma(-j-\bar{m})}{\Gamma(-2j-1)
1905: \Gamma(j+1+m)\Gamma(j+1-\bar{m})}, \cr
1906: \nu(k) &\equiv \frac{1}{\pi}\frac{\Gamma(1-\frac{1}{k})}
1907: {\Gamma(1+\frac{1}{k})}~, \label{strt}
1908: \end{align}
1909: The effects of the winding tachyon condensation can be seen in the $1/k$ suppressed factor in the exact formula as $\frac{\Gamma(1+\frac{2j+1}{k})}{\Gamma(1-\frac{2j+1}{k})}$. As is well-known in Liouville field theory, the poles in the correlation function appear when the screening interaction coming from the $\mathcal{N}=2$ superpotential $W= \mu e^{\frac{1}{Q}\Phi}$ satisfies the screening condition for the Liouville momenta $\phi$. Indeed, the perturbative Liouville insertion predicts poles in the two-point functions exactly as indicated by the factor $\frac{\Gamma(1+\frac{2j+1}{k})}{\Gamma(1-\frac{2j+1}{k})}$.
1910: 
1911: Another important aspect of the FZZ duality is that it has provided a perspective on the winding number non-conservation process. In the $SL(2;\br)^{(A)}/U(1)$ axial coset model, one can define an asymptotic winding quantum number by $\omega$. However, since the cigar geometry has a trivial fundamental group, the winding number is not a conserved quantity. In the free field construction of the $SL(2;\br)^{(A)}/U(1)$ coset model (such as the one based on the Wakimoto construction of the $SL(2;\br)$ current algebra), it is difficult to compute the winding number violating correlation functions. Indeed this was the first motivation of FZZ to propose the dual description.\footnote{At the same time, FZZ has also given an ingenious way to compute the winding violating correlation function within the $SL(2;\br)/U(1)$ coset model by introducing the dual operators.} 
1912: 
1913: Situations are worse in the naive T-dualized trumpet geometry. In the trumpet metric, it appears that we have a $U(1)$ isometry along $\tilde{\theta}$ that is the dual coordinate for $\theta$, suggesting that the momentum quantum number as well as the winding quantum number are well-defined conserved quantities. The breaking of the winding number (or momentum mode in the T-dual picture) is quite obscure: the origin of the winding non-conservation, i.e. the fixed point of the $U(1)$ action, has now become the singularity of the target space.\footnote{It is well-known that when we gauge the axial symmetry, the vector current has an anomaly and vice versa, and this is indeed the origin of this apparent paradox. In the same token, the $U(1)$ isometry of the vector coset is broken down to $\bz_k$. We should be, therefore, careful when we talk about the ``T-duality" of the trumpet.}
1914: 
1915: The resolution of this puzzle is given by the FZZ duality. In the T-dualized picture, the singularity is removed by the tachyon condensation, or $\mathcal{N}=2$ Liouville superpotential. At the same time, the $\mathcal{N}=2$ superpotential explicitly breaks the translation invariance along the $\theta$ direction, which gives the origin of the momentum non-conservation in the T-dual picture. Actually, the explicit breaking of the momentum conservation is quite useful to compute the winding number violating process in $SL(2;\br)/U(1)$ coset model: by a direct insertion of the $\mathcal{N}=2$ Liouville superpotential, the winding number violating process can be computed perturbatively.
1916: 
1917: We end this section with three remarks
1918: \begin{itemize}
1919: 	\item Supersymmetric $SL(2;\br)/U(1)$ coset model has a conserved $U(1)_R$ current. By taking quotient of the theory with this $U(1)_R$  current, we obtain the duality between the bosonic $SL(2;\br)/U(1)$ coset model and the sine-Liouville theory \cite{Karczmarek:2004bw}. The sine-Liouville theory has the potential
1920: \begin{align}
1921: V(\phi,Y) &= \mu (S^+ + S^-) \cr
1922: S^{\pm} &=  e^{-\frac{1}{\cQ}(\phi \pm \sqrt{1+\cQ^2}iY)}
1923: \equiv e^{-\sqrt{\frac{\kappa-2}{2}} \phi \mp
1924: \sqrt{\frac{\kappa}{2}} iY}~,
1925: ~~~ (\cQ=\sqrt{2/(\kappa-2)})~.
1926: \end{align}
1927: We note that the potential preserves the $W_{\infty}$ symmetry as a side remark, which makes the model integrable \cite{Baseilhac:1998eq,Lukyanov:2003nj}. In their original work (FZZ), they proposed the duality between the bosonic $SL(2;\br)/U(1)$ coset model and the bosonic sine-Liouville theory.
1928: 
1929: 	\item There is a small controversial issue in the interpretation of the FZZ duality. Our standpoint has been that the dual description of the cigar geometry is given by the $\mathcal{N}=2 $ Liouville theory, and the $\mathcal{N}=2$ Liouville superpotential does not appear in the original cigar geometry explicitly (otherwise the source of the winding number non-conservation is two-fold). The other common interpretation of the FZZ duality is that the winding tachyon condensation ($\mathcal{N}=2$ Liouville superpotential written in the dual coordinate) also appears in the original cigar geometry. This interpretation is natural in the sense that it gives a natural explanation about the coexistence of the poles coming from the geometry part and the Liouville insertion part. Whichever interpretation one may take, we believe that what we call the supersymmetric $SL(2;\br)/U(1)$ theory and the $\mathcal{N}=2$ Liouville theory is identical, and the structure constant, e.g. the two-point function is uniquely given by formulae like \eqref{strt}.
1930: 
1931: 
1932: 	\item So far, we have focused on the Euclidean $SL(2;\br)/U(1)$ coset model. However, things are unclear in the Lorentzian $SL(2;\br)/U(1)$ coset model, where the dual $\mathcal{N}=2$ Liouville theory is unavailable. A naive analytic continuation of the $\mathcal{N}=2$ superpotential gives a wrong Liouville wall, which is localized near the weakly coupled region \cite{Hikida:2004mp}. Furthermore, the Lorentzian coset does not have a winding mode, so the interpretation of the winding tachyon condensation is not evident. Nevertheless, we believe that the exact structure constants are given by the analytic continuation of the exact results for the Euclidean coset since the analytic continuation correctly reproduces the mini-superspace part. The clear explanation of the origin of the extra poles in the Lorentzian coset is still an open question.\footnote{The origin might be given by the degrees of freedom near the horizon on which we mentioned in section \ref{sec:3-2-3}. A related interpretation based on the idea of stretched horizon has been given in \cite{Kutasov:2005rr}.}
1933: \end{itemize}
1934: 
1935: \newpage
1936: \sectiono{Black Hole - String Transition}\label{sec:4}
1937: In this section, we review ``black hole - string transition". The transition is believed to be a fundamental property of the quantum black hole in the non-BPS regime. We will also see that the transition is related to the thermal winding tachyon condensation.
1938: 
1939: The organization of the section is as follows. In section \ref{sec:4-1}, we formulate the ``black hole - string transition" in general dimensions. In section \ref{sec:4-2}, we specialize in the two-dimensional case, where $\alpha'$ exact treatment is possible. In section \ref{sec:4-3}, we briefly summarize the current status of the black hole - string transition in other solvable backgrounds.
1940: 
1941: 
1942: 
1943: \subsection{In general dimensions}\label{sec:4-1}
1944: 
1945: One of the most profound results in (semi-)classical gravity is the thermodynamics of the black hole. Thus one of the most significant benchmarks of any theory of quantum gravity is to provide a satisfactory understanding of the thermodynamics of the black hole. Especially, understanding of the black hole entropy from the microscopic viewpoint has been one of the greatest achievements of the string theory as a quantum theory of gravity \cite{Strominger:1996sh}. 
1946: 
1947: Let us consider the Schwarzshild black hole in the string theory as a simplest example of the non-extremal black hole system.\footnote{The exact quantization of the string in the Schwarzshild black hole is not known. However, since one can make the curvature of the Schwarzshild black hole arbitrarily small outside the horizon, it is natural to assume the existence of string solutions asymptotically given by the Schwarzshild black hole. The existence of the $SL(2;\br)/U(1)$ two-dimensional black hole strongly supports this assumption.} 
1948: %The Schwarzshild black hole is the unique solution of the vacuum Einstein equation with spherical symmetry.
1949: The Schwarzshild black hole in any dimension is completely determined by the parameter $r_h$ that determines the horizon size. When $r_h \gg l_s $, the classical supergravity description is good (at least outside of the horizon), and we can trust the effective supergravity action to discuss the properties of the black hole. 
1950: 
1951: If one gradually decreases the horizon size $r_h$, the effects of higher derivative corrections coming from the underlying quantum gravity will become important. Within the superstring theory, some higher derivative corrections are known, and it has been shown that these corrections will beautifully explain the apparent mismatch between the macroscopic derivation of the small charge BPS black hole entropy and the microscopic derivation from the string theory (see e.g. \cite{Dabholkar:2004yr}). In the non-extremal cases we are discussing now, we have not yet completely grasped the structure of the higher derivative corrections and the quantitative match of the black hole entropy, but the guiding principle is summarized by the so-called ``black hole - string transition" or ``black hole - string crossover" introduced in \cite{'tHooft:1987tz,Holzhey:1991bx,Susskind:1993ws,Horowitz:1996nw,Sen:1995in}.
1952: 
1953: When $r_h \le l_s$, the geometrical description of the black hole breaks down and it should be replaced with the microscopic description based on the quantum strings. This is natural because the string theory has a natural cutoff  given by the string length $l_s$ as a length scale, and the objects smaller than $l_s$ do not possess an ordinary geometrical meaning. The principle of the ``black hole - string transition" is that the black hole can be understood either as the higher excitation of the strings or as the classical solution of the (higher derivative) gravities. Especially, the crossover is parametrically smooth as a function of the coupling constant $g_s$ and $l_s$. 
1954: 
1955: In the Schwarzshild black hole example, we can roughly estimate the transition point and the ``black hole - string crossover" as follows. Let us assume the four dimensional Schwarzshild black hole for definiteness. The four-dimensional Newton constant $G$ is given by $G \sim g_s^2 l_s^2$, so the Schwarzshild radius of the string is estimated as $r_{0} = m_{\rm str} G \sim m_{\rm str} g_s^2 l_s^2$ with the mass of excited string $m^2_{\rm str} \sim \frac{n}{l_s^2}$, where $n$ denotes the oscillator level. At the black hole - string transition point $r_0 \sim l_s$, we have
1956: \begin{align}
1957: \frac{l_s^2}{G} \sim \frac{1}{g_s^2} \sim \sqrt{n} \ .
1958: \end{align}
1959: Thus, the classical Bekenstein entropy is given by $S_{Bek} \sim \frac{r_0^2}{G} \sim \frac{l_s^2}{G} \sim \sqrt{n}$, which indeed agrees with the entropy of the perturbative string expected from the Cardy formula up to a numerical factor. Alternatively speaking, one can say that the requirement of the smooth overlap of the entropy demands that $r_0 \sim l_s$ should be the ``black hole - string transition" point.
1960: 
1961: Another important concept associated with the $\alpha'$ corrections to the geometry is the stretched horizon \cite{Susskind:1993aa,Susskind:1993ws,Kutasov:2005rr}. We can formulate the stretched horizon based on the local temperature of the geometry. As is the case with the two-dimensional black hole, any neutral black hole has an intrinsic temperature determined by the periodicity of the Euclidean time (Hawking temperature). From the Lorentzian viewpoint, the temperature is defined by the observer at spacial infinity. From an observer at a fixed proper distance $R$ from the horizon, the Hawking radiation is observed with much higher temperature
1962: \begin{align}
1963: T_{u}(R) = \frac{T_{\rm Hw}}{\sqrt{g_{00}(R)}} \ , \label{uhw}
1964: \end{align}
1965: due to the gravitational red-shift. On the other hand, the string theory has the ``highest temperature" determined by the Hagedorn temperature. Since the number of perturbative string states grows exponentially as a function of energy (mass): \begin{align}
1966: Z(\beta) = \mathrm{Tr} e^{-\beta E} \sim \int \dd M \rho(M) e^{-\beta M} \ . \label{hgr}
1967: \end{align}
1968: with the density of states given by $\rho(M) \sim e^{\beta_{\rm Hg} M} $, the partition function of the perturbative string theory is ill-defined beyond the Hagedorn temperature $\beta < \beta_{\rm Hg}$. There we expect that the string interactions are much more important and the strings will disentangle. 
1969: 
1970: Now let us return to the Hawking radiation. From \eqref{uhw}, one can see that the (red-shifted) temperature becomes infinite at the classical horizon. Actually, before reaching the event horizon  we will encounter the radius when the local temperature exceeds the Hagedorn temperature. The local Hagedorn transition blurs the local geometry near the black hole horizon. This is what we call the stretched horizon. Note that we can make the curvature at the horizon arbitrarily small, and in this regime, the size of the stretched horizon is of order one in the string unit.
1971: 
1972: It is interesting to consider some extreme limits of the above discussions. The first example is the large $T_{\rm Hw}$ limit: what happens if the Hawking temperature in the asymptotic infinity is larger than $T_{\rm Hg}$? We expect that the stretched horizon completely blur the black hole geometry. Indeed in the leading order estimation of the four-dimensional Schwarzshild black hole, we have $\beta_{\rm Hw} = r_0$ and  $\beta_{\rm Hg} = \text{const}$, and such scenario occurs when $r_0 \sim l_s$. It is interesting to note that the condition roughly coincides with that for the ``black hole - string transition". In section \ref{sec:4-2}, we will see this coincidence is exact (after taking $\alpha'$ corrections into account) in the two-dimensional black hole that is an exactly solvable string background. 
1973: 
1974: Another limit is the (extremal) charged black hole solution. In the charged black hole examples, the above discussion based on the Hagedorn temperature and the Hawking temperature should be generalized. This is because, as pointed out in \cite{Horowitz:1996nw},  we can arbitrarily lower the Hawking temperature while keeping possible $\alpha'$ corrections large. In other worlds, one can make the transition temperature arbitrarily lower than the Hagedorn temperature. The most extreme case is the (BPS) extremal black hole, where the Hawking temperature is zero. The generalization proposed in \cite{Giveon:2005jv} states that the ``black hole - string transition" occurs when the Hawking temperature coincides with the temperature of the free-string with the same mass {\it and} charge. We will briefly review their discussions later in section \ref{sec:4-3}.
1975: 
1976: It is also instructive to recapitulate the problem from the Euclidean approach. In the flat Minkowski space, the Hagedorn divergence of the partition function can be attributed to the thermal winding tachyon condensation \cite{Polchinski:1985zf,Sathiapalan:1986db,Kogan:1987jd,O'Brien:1987pn,Atick:1988si}. 
1977: We begin with the more precise version of \eqref{hgr}.
1978: \begin{align}
1979: \beta F &= \mathrm{Tr}_{\mathrm{phys}} \log (1- e^{-\beta E}) \cr
1980:     &= \int_{-\infty}^{\infty}\dd\tau_2 \int_{-1/2}^{1/2} \dd\tau_1 \frac{1}{\tau_2} \mathrm{Tr}_{\mathrm{CFT}} q^{L_0}q^{\bar{L}_0} \ . \label{hgg}
1981: \end{align}
1982: In the second line, we have introduced the Schwinger parameter $\tau_2$ and the level matching condition by 
1983: \begin{align}
1984: \int_{-1/2}^{1/2} \dd\tau_1 e^{2\pi i \tau_1 (L_0-\bar{L}_0)} \ .
1985: \end{align}
1986: The trace is taken over the original space-like CFT with an additional free $\mathbb{S}_1$ CFT whose radius is $\beta$ {\it restricted to the momentum mode}. The Hagedorn divergence appears in the ultraviolet region $\tau_2 \to 0$. Now let us use the Polchinski's trick \cite{Polchinski:1985zf} to rewrite the thermal partition function \eqref{hgg} as the string 1-loop partition function
1987: \begin{align}
1988: \beta F &= \int_{\mathcal{F}}\frac{\dd\tau^2}{\tau_2} \mathrm{Tr}_{\mathrm{CFT}\times \mathbb{S}^1} q^{L_0}q^{\bar{L_0}} \ ,
1989: \end{align}
1990: where $\mathcal{F}$ is the fundamental domain of the torus, and the trace is taken over the original CFT with the free $\mathbb{S}^1$ CFT {\it including winding modes}. The Hagedorn divergence is now translated to the IR instability $\tau_2 \to \infty$. Apart from the ground state tachyon that should be GSO-projected out in the supersymmetric theory, a possible instability comes from the thermal winding tachyon whose mass is given by
1991: \begin{align}
1992: m(\beta)^2 = -1 + \beta^2 \ .
1993: \end{align}
1994: When $m(\beta)^2 < 0$, the Hagedorn instability occurs. In this way, we can understand the Hagedorn divergence as the appearance of the winding tachyon in the Euclideanized thermal string theory.
1995: 
1996: The argument above suggests that when the thermal direction shrinks enough to admit ``winding tachyon" in the Euclidean spectrum, the Hagedorn phase transition occurs. Assuming a semiclassical quantization of string in the Schwarzshild black hole, a similar situation occurs in the thermal string theory in the Euclidean Schwarzshild black hole background. The thermal winding tachyon has an effective mass
1997: \begin{align}
1998: m^2(r) = -1 + r_0^2\left(1-\frac{r_0}{r}\right) \ .
1999: \end{align}
2000: At the point where $m^2(r)$ becomes negative, the black hole develops a stretched horizon, and when $m^2(\infty) < 0$, we expect the ``black hole - string" phase transition. We will see later that the winding tachyon is crucial in the two-dimensional black hole and its exact ``black hole - string phase transition".
2001: 
2002: Recently Horowitz \cite{Horowitz:2005vp} studied the real-time winding tachyon condensation in the black hole system. If one considers a compactified black string solution, the extra dimension can show a winding tachyon condensation as the direction shrinks toward the black hole singularity. After the winding tachyon condensation, the black hole evaporates as a bubble of nothing. This process is proposed to be a new interesting end point of the Hawking black hole evaporation (see \cite{Ross:2005ms,Bergman:2005qf,Horowitz:2006mr,Dine:2006we} for related studies). The winding tachyon condensation could also give a solution of the cosmological singularity problems as studied in \cite{McGreevy:2005ci,Nakayama:2006gt}.
2003: 
2004: \subsection{Two-dimensional black hole case}\label{sec:4-2}
2005: To discuss the ``black hole - string transition" introduced in section \ref{sec:4-1} in a more quantitative manner, it is imperative to study the exact string background rather than the approximate Schwarzshild black hole solution. Especially, the arguments related to the (thermal) winding tachyon condensation is rather speculative, and a demonstration based on the exactly solvable string background would be highly desirable. As we have seen in section \ref{sec:2}, the simplest exactly solvable (non-BPS) black hole background is the two-dimensional black hole. In this subsection, we specialize in the ``black hole - string transition" in the two-dimensional black hole.\footnote{Of course, what we mean by the ``two-dimensional black hole" includes the embedding into the superstring theory such as the black NS5-brane background, so our results have a direct application to the ten-dimensional critical string theories.}
2006: 
2007: Let us consider the supersymmetric $SL(2;\br)/U(1)$ coset model. As we discussed in section \ref{sec:2-4}, the two-dimensional black hole has the Hawking temperature
2008: \begin{align}
2009: T_{\rm Hw} = \frac{1}{\beta_{\rm Hw}} = \frac{1}{2\pi\sqrt{\alpha'k}} \ .
2010: \end{align}
2011: Since the two-dimensional black hole is asymptotically a linear dilaton theory,  the Hagedorn temperature shows a $1/k$ corrected shift compared with the flat Minkowski theory\footnote{When we mention the Hagedorn temperature of the two-dimensional black hole, we always assume that the criticality condition of the string theory is satisfied by adding {\it non-dilatonic} CFTs. The NS5-brane background is a typical example.}:
2012: \begin{align}
2013: T_{\rm Hg} = \frac{1}{\beta_{\rm Hg}} = \frac{1}{4\pi\sqrt{1-\frac{1}{2k}}} \ . \label{Hg}
2014: \end{align}
2015: To derive this formula, one should first note that the $SL(2;\br)/U(1)$ coset model has a mismatch between the genuine central charge $c^{SL(2;\br)/U(1)} =  3 + \frac{6}{k}$ and the effective central charge $c_{\mathrm{eff}}^{SL(2;\br)/U(1)} = 3$ due to the asymptotic linear dilaton.\footnote{We can also see this directly from the one-loop partition function and the spectrum. See section \ref{sec:3} and appendix A.} Therefore, the total theory has a deficit effective central charge $c_{\mathrm{eff}}^{\mathrm total} = 12 - \frac{6}{k}$ after the subtraction of the ghost contribution. Now we recall the Cardy formula:
2016: \begin{align}
2017: \rho(M) \sim \exp\left(2\pi\sqrt{\frac{c_{\mathrm{eff}}}{12}}M + 2\pi \sqrt{\frac{\bar{c}_{\mathrm {eff}}}{12}}M \right) \ ,
2018: \end{align}
2019: which immediately gives the Hagedorn temperature \eqref{Hg}. 
2020: 
2021: For later references, we present here a similar formula for the bosonic two-dimensional black hole. The Hawking temperature and the Hagedorn temperature is given by
2022: \begin{align}
2023: T_{\rm Hw} = \frac{1}{\beta_{\rm Hw}} = \frac{1}{2\pi\sqrt{\alpha'\kappa}} \ . \label{hwb}
2024: \end{align}
2025: and
2026: \begin{align}
2027: T_{\rm Hg} =\frac{1}{\beta_{\rm Hg}} = \frac{1}{4\pi\sqrt{2-\frac{1}{2(\kappa-2)}}} \ . \label{Hgb}
2028: \end{align}
2029: There is no apparent reason to exclude possible $\alpha'$ corrections to the Hawking temperature in the bosonic string theory, but the exact string quantization reveals that the formula \eqref{hwb} is the correct one.\footnote{In general, the Hawking temperature is classically determined solely from the information near the event horizon (the Rindler limit), where the curvature and the $\alpha'$ corrections could become large. The effects of such $\alpha'$ corrections and possible renormalization of the Hawking temperature are interesting subjects to study.} We will return to this problem when we discuss the exact boundary states of the probe rolling D-brane in section \ref{sec:8}. On the other hand, the Hagedorn temperature here is obtained from the exact string quantization and trustful.
2030: 
2031: From the general discussions in section \ref{sec:4-1}, we expect ``black hole -  string" at $k=1$ (or $\kappa=3$ for the bosonic case) when the Hawking temperature and the Hagedorn temperature coincide. At this point, the stretched horizon becomes so large that it will swallow the complete space-time. This ``black hole - string transition" in the two dimension black hole is induced by the strong $\alpha'$ corrections: when $k$ is large (recall $1/k$ correction corresponds to $\alpha'$ correction) the Hawking temperature is much larger than the Hagedorn temperature, and the geometry is not disturbed by the back-reaction of the Hawking radiation. When $k$ becomes smaller, $1/k$ corrections will become more and more important, and at the phase transition point, i.e. at $k=1$, physics changes drastically. One of the main focus of this thesis is to study this transition from the rolling D-brane probe.
2032: 
2033: At this point, we would like to point out that the ``black hole - string" phase transition of the two-dimensional black hole does not involve the string coupling $g_s$ in the discussion. This is one of the features of the two-dimensional black hole that we can clearly separate the (typically more difficult) problem of the genus expansion from the more tractable $\alpha'$ corrections in order to understand the ``black hole - string transition".
2034: 
2035: What is the origin of the strong $1/k$ correction? As we have mentioned earlier in section \ref{sec:3-1}, the metric for the supersymmetric two-dimensional black hole does not receive perturbative $1/k$ corrections. The origin of the (nonperturbative) $1/k$ corrections that trigger the ``black hole - string transition" is most clearly seen in the Wick rotated Euclidean two-dimensional black hole for which the dual description is available. 
2036: 
2037: In the dual description, the two-dimensional black hole is described by the $\mathcal{N}=2$ Liouville theory. The $\mathcal{N}=2$ superpotential
2038: \begin{align}
2039: W(\Phi) = \mu \int d^2\theta e^{\frac{1}{Q}\Phi} \label{Liousup}
2040: \end{align}
2041: can be seen as the localized (winding) tachyon condensation.\footnote{The duality between the $SL(2;\br)^{(A)}/U(1)$ coset and the $\mathcal{N}=2$ Liouville theory is a kind of T-duality as we discussed in section \ref{sec:3-4}. Thus the condensation of the momentum mode in $\mathcal{N}=2$ Liouville theory can be regarded as the condensation of the winding mode in the original $SL(2;\br)^{(A)}/U(1)$ coset model.} The condensation is a localized mode because the Liouville momentum $j$ corresponding to the superpotential \eqref{Liousup} is not given by the continuous series $j=-\frac{1}{2}+ip$, but lies in the discrete series.
2042: 
2043: The crucial observation has been already made early in \cite{Kutasov:1990ua} in the context of the noncritical superstring theory. The spirit is close to the discussions given in section \ref{sec:2-3} and \ref{sec:3-4}. The superpotential \label{Liousup} is a normalizable perturbation if $\frac{1}{Q}>\frac{Q}{2}$ (i.e. $k >1$) and it is a non-normalizable deformation otherwise. In the language of the noncritical string theory, the $\mathcal{N}=2$ super Liouville potential satisfies the Seiberg bound \cite{Seiberg:1990eb} only when  $\frac{1}{Q}<\frac{Q}{2}$  holds. This directly means that the $\mathcal{N}=2$ Liouville description is good for $k<1$ and the two-dimensional black hole description is good for $k>1$. The transition point is exactly at $k=1$.\footnote
2044:    {Another interesting observation related
2045:     to the $k=1$ transition is the following. If we consider
2046:         a two-dimensional $U(1)$ gauge theory in the ultraviolet that
2047:         flows to $SL(2;\br)/U(1)$ coset theory in the infrared (as was
2048:         introduced in \cite{Hori:2001ax} to prove the mirror duality to the
2049:         $\cN=2$ Liouville theory),
2050:   the central charge of the $U(1)$ gauge theory
2051:   is given by $9$. Since the IR $SL(2;\br)/U(1)$ coset theory
2052:   has a central charge $c=3(1+\frac{2}{k})$, there is an
2053:   apparent contradiction to Zamolodchikov's $c$-theorem
2054:   if the level $k<1$ is considered. However, we should note that
2055:   $SL(2;\br)/U(1)$ coset theory is dilatonic so that the effective
2056:   central charge is always given by $3$.}
2057: 
2058: We can repeat the same analysis for the bosonic $SL(2;\br)/U(1)$ coset. The duality between the bosonic $SL(2;\br)/U(1)$ and the sine-Liouville theory, together with the Seiberg bound, leads to the conclusion that $\kappa=3$ is the phase transition point. The potential is given by
2059: \begin{align}
2060: V = \mu (S^+ + S^-) \  , \ \ \ S^{\pm} =  e^{-\frac{1}{\cQ}(\phi \pm \sqrt{1+\cQ^2}iY)}
2061: \equiv e^{-\sqrt{\frac{\kappa-2}{2}} \phi \mp
2062: \sqrt{\frac{\kappa}{2}} iY}~,
2063: ~~~ (\cQ=\sqrt{2/(\kappa-2)})~,
2064: \end{align}
2065: and the normalizability changes precisely at $\kappa = 3$.
2066: Assuming that this occurs when the Hawking temperature and the Hagedorn temperature coincides, we have verified that the Hawking temperature of the bosonic two-dimensional black hole is not renormalized. We will see another support from the probe rolling D-brane later in section \ref{sec:8}.
2067: 
2068: We could argue this transition without using the dual Liouville picture \cite{Karczmarek:2004bw}. The black hole perturbation descends from the $SL(2;\br)$ states
2069: \begin{align}
2070: J_{-1}^+ \bar{J}_{-1}^+|j=-1; m = \bar{m} = -1 \rangle \ .
2071: \end{align}
2072: The normalizability of such states (see section \ref{sec:3-2-2}) demand
2073: \begin{align}
2074: -\frac{1+k}{2} < j  < -\frac{1}{2} \ 
2075: \end{align}
2076: with $j=-1$, which suggests the same phase transition point $k=1$ (or $\kappa=3$).
2077: 
2078: The situation in the Lorentzian two-dimensional black hole is less clear. We cannot perform the Wick rotation to the winding tachyon potential \eqref{Liousup} naively because the time is continuous and there is no apparent winding mode in the Lorentzian two-dimensional black hole. The same thing can be said in the Hagedorn instability of the free string theory in the flat Minkowski space: the existence of the thermal winding tachyon in the Wick rotated theory does not mean the tachyonic instability in the real time physics. Rather it should be understood as the phase transition associated with the thermal dissolution of strings. At the temperature beyond the Hagedorn point, there would be no distinction between the gas of strings and the black hole.
2079: 
2080: 
2081: 
2082: \subsection{Other solvable backgrounds}\label{sec:4-3}
2083: There are many other exactly solvable string theory backgrounds that exhibit the ``string black hole transition". Most of the examples are more or less related to the $SL(2;\br)$ WZNW model. In this subsection, we will briefly review the transition in such backgrounds.
2084: 
2085: The black hole - string transition across $k=1$ also has a natural
2086: interpretation in terms of the holographic principle, as recently
2087: discussed in \cite{Giveon:2005mi}. Adding $Q_1$ fundamental strings to $k$
2088: NS5-branes (more generally Calabi-Yau singularities) as we reviewed in section 2.4, one obtains the familiar bulk geometry of the
2089: $AdS_3/CFT_2$-duality. In this context, the density of states of the
2090: dual conformal field theory is given by the naive Cardy formula
2091: $S=2\pi\sqrt{\frac{cL_0}{6}}+2\pi\sqrt{\frac{\bar{c}\bar{L}_0}{6}}$
2092: with $c = 6 k Q_1$ for $k>1$, but not for $k<1$. Rather, the central
2093: charge that should be used in the Cardy formula is replaced by an
2094: effective one $c_{\rm eff}= 6Q_1(2-\frac{1}{k})$ \cite{Kutasov:1990ua}.
2095: 
2096: The origin of the difference between the $c_{\mathrm{eff}}$ and $c$ is again the normalizability of a certain operator. The $SL(2;\mathbb{C})$ vacuum of the dual CFT corresponds to the states
2097: \begin{align}
2098: J_{-1}^+ \bar{J}_{-1}^+|j=-1; m = \bar{m} = -1 \rangle \ 
2099: \end{align}
2100: in the world-sheet $SL(2;\br)$ WZNW model, and as we have seen several times, for $k>1$, the operator is normalizable, and $c_{\mathrm{eff}} = c$. On the other hand, for $k<1$, the operator is non-normalizable, and we expect $c_{\mathrm{eff}} <c$. A short computation based on the string description gives $c_{\rm eff}= 6Q_1(2-\frac{1}{k})$.
2101: 
2102: We note that for $k>1$, the BTZ black hole excitation is normalizable and the partition function and the entropy is dominated by the Bekenstein-Hawking entropy of the BTZ black hole while for $k<1$, the BTZ black hole excitation is non-normalizable and the entropy is solely explained by the string excitations. This argument is completely in agreement with the ``black hole - string transition" picture at $k=1$.
2103: 
2104: Another interesting generalization is the two-dimensional charged black hole. We consider the asymmetric coset
2105: \begin{align}
2106: \frac{SL(2;\br)_k\times U(1)_L}{U(1)} \ , \label{chtbk}
2107: \end{align}
2108: where the $U(1)$ gauging acts on one of the (space-like) left-moving current in $SL(2;\br)$ and a linear combination of the right-moving current of $SL(2;\br)$ and $U(1)_L$. After the Kaluza-Klein reduction, the geometry of \eqref{chtbk} is described by the metric ($Q^2 = \frac{2}{k}$)
2109: \begin{align}
2110: \dd s^2 = \dd\phi^2 - \left(\frac{\tanh\frac{Q}{2}\phi}{1-a^2\tanh^2\frac{Q}{2}\phi}\right)^2 \dd\theta^2 \ ,
2111: \end{align}
2112: the dilaton
2113: \begin{align}
2114: \Phi = \Phi_0 - \frac{1}{2}\log\left(1+(1-a^2)\sinh^2\frac{Q}{2}\phi\right) \ ,
2115: \end{align}
2116: and the gauge field
2117: \begin{align}
2118: A = \frac{a\tanh^2\frac{Q}{2}\phi}{1-a^2\tanh^2\frac{Q}{2}\phi} \dd\theta \ .
2119: \end{align}
2120: Here $a^2$ is related to the mass $m$ and the charge $q$ of the black hole as
2121: \begin{align}
2122: a^2 = \frac{m-\sqrt{m^2-q^2}}{m+\sqrt{m^2-q^2}} \ .
2123: \end{align}
2124: At $a = 0$, the model reduces to the undeformed $SL(2;\br)/U(1)$ black hole (and a compact boson).
2125: 
2126: The Hawking temperature of the black hole (e.g. from the Euclidean geometry) is given by
2127: \begin{align}
2128: \beta_{\rm Hw} = \frac{4\pi}{Q} \frac{1}{1-a^2} \ .
2129: \end{align}
2130: On the other hand, the Hagedorn temperature is given by 
2131: \begin{align}
2132: T_{\rm Hg} =\frac{1}{\beta_{\rm Hg}} = \frac{1}{4\pi\sqrt{1-\frac{1}{2k}}} \ , 
2133: \end{align}
2134: irrespective of the deformation parameter $a$.
2135: 
2136: From the world-sheet perspective, the ``black hole - string transition" of the charged two-dimensional black hole inherits from the $SL(2;\br)$ WZNW model and the transition point should be $k=1$. This is different from the naive guess based on the relation $\beta_{\rm Hw} = \beta_{\rm Hg}$. A resolution proposed in \cite{Giveon:2005jv} is that the more precise definition of the transition temperature is when the Hawking temperature coincides with the temperature of the string that has the same mass and charge of the black hole. 
2137: 
2138: In this example, the entropy of the string with charge $q$ is given by\footnote{The shift is due to $q$-amount of right-moving $U(1)$ charge: we are summing over the string states with fixed $U(1)$ charge $q$ instead of summing over all states.}
2139: \begin{align}
2140: S = 2\pi \sqrt{1-\frac{Q^2}{4}} \left(m+\sqrt{m^2-q^2}\right)
2141: \end{align}
2142: resulting in the corresponding string temperature
2143: \begin{align}
2144: \beta_{\rm str} = \left.\frac{\partial S}{\partial m} \right|_{q} = \sqrt{1-\frac{Q^2}{4}}\frac{4\pi}{1-a^2} \ .
2145: \end{align}
2146: It is easy to see that the condition $\beta_{\rm str} = \beta_{\rm Hw}$ exactly reproduces the CFT computation, i.e. $k=1$.
2147: \newpage
2148: \sectiono{Tachyon Radion Correspondence}\label{sec:5}
2149: In this section, we review the tachyon radion correspondence, which is one of the greatest motivations to study the rolling D-brane in the two-dimensional black hole system. The correspondence says that the dynamics of the open-string tachyon condensation may be geometrically realized by the rolling D-brane system. 
2150: 
2151: The organization of the section is as follows. In section \ref{sec:5-1}, we overview the rolling tachyon problem. In section \ref{sec:5-2}, we study the closed string emission rate from the rolling tachyon boundary states and their variations.\footnote{This part of the thesis is based on \cite{Nakayama:2006qm}.} In section \ref{sec:5-3}, we study the correspondence at the classical level. In section \ref{sec:5-4}, we summarize our results on the quantum correspondence. In section \ref{sec:5-5} some cosmological implications are studied.
2152: 
2153: 
2154: \subsection{Rolling tachyon}\label{sec:5-1}
2155: \subsubsection{overview}\label{sec:5-1-1}
2156: In the days of early developments of string theory, tachyon used to be thought of as a nuisance in constructing realistic models for particle physics of our world. In recent years, open-string tachyons have obtained civil rights and have played more and more important roles in acquiring our knowledge on the nonperturbative D-brane physics with (spontaneously) broken SUSY. In addition, they have been even providing phenomenological applications such as brane inflation. More recently, the closed string tachyons (especially localized winding tachyon) have attracted much attention in relation to the topological change \cite{Adams:2001sv,Adams:2005rb} and the resolution of singularities \cite{McGreevy:2005ci}.
2157: 
2158: One of the important steps in understanding the physics of unstable D-branes is Sen's conjecture with subsequent advancement (see \cite{Sen:2004nf} for a review), which states that the decaying process of the unstable D-branes can be regarded as the open string tachyon condensation. In particular, the energy difference between the false (perturbative) tachyonic vacuum and the true vacuum of the open string tachyon potential should explain the tension of the decaying D-brane exactly. Furthermore, the cohomology of the open string theory at the true vacuum must vanish. In the context of the open string field theory, these conjectures have been analytically proved in \cite{Schnabl:2005gv,Ellwood:2006ba}. 
2159: 
2160: More interesting aspects of the tachyon dynamics is to study its time evolution \cite{Sen:2002nu,Sen:2002in,Sen:2002an}. Based on the effective field theory analysis (which has been confirmed by the exact boundary states analysis later), it was found that the late time evolution of the open-string tachyon gives rise to the so-called ``tachyon matter", which is a pressureless fluid. Such a ``rolling tachyon" evolution has provided us with novel understanding of the tachyon condensation and time-dependent physics in string theory. The feasibility to construct the exact boundary states enables us to study the highly non-supersymmetric time evolution in a quantitative way.
2161: 
2162: Let us begin with the effective DBI type action for the rolling tachyon
2163: \begin{align}
2164: S = - \int \dd t V(T)\sqrt{1-\dot{T}^2} \ . \label{tDBI}
2165: \end{align}
2166: Since we will focus on the homogeneous decay, we have assumed the D0-brane action without loss of generality. The effective potential $V(T)$ takes the form
2167: \begin{align}
2168: V(T) = M_0 \frac{1}{\cosh\frac{T}{2x}} \ ,
2169: \end{align}
2170: where the D0-brane tension $M_0 \propto \frac{1}{g_s} $. For the non-BPS D-branes in supersymmetric theory, $x=1$, and for the unstable D-branes in bosonic string theory, $x=1/2$.
2171: The solution of the equation of motion is given by
2172: \begin{align}
2173: \sinh\frac{T}{2x} = a\cosh\frac{t}{2x} \ , \label{seom}
2174: \end{align}
2175: leading to the classical energy momentum tensor:
2176: \begin{align}
2177: T_{\mu\nu} = \frac{V(T)\partial_\mu T\partial_\nu T}{\sqrt{1+\eta^{\mu\nu}\partial_\mu T\partial_\nu T}} - V(T) \eta_{\mu\nu} \sqrt{1+\eta^{\mu\nu}\partial_\mu T\partial_\nu T} \ .
2178: \end{align}
2179: We are interested in the late time behavior of the energy momentum tensor, which is explicitly given by 
2180: \begin{align}
2181: T_{00} & \sim E \cr
2182: T_{ij} & \sim -E \exp(-t/x) \delta_{ij} \ ,
2183: \end{align}
2184: where $E = M_0/\sqrt{1+a^2}$. As we mentioned before, we have obtained the pressureless dust as a final product of the D0-brane decay.
2185: 
2186: The energy momentum tensor yields the coupling of the rolling D-brane to the gravity. In order to study the coupling to higher string modes, we need the exact boundary state that describes the rolling D-brane. In the boundary conformal field theory approach, we introduce the boundary interaction\footnote{We focus on the bosonic case for simplicity. The generalization to the non-BPS D-branes in superstring theory is straightforward.}
2187: \begin{align}
2188: \delta S_{\mathrm{full}} = \tilde{\lambda} \int \dd s \cosh X^0(s) \ , \label{fulli}
2189: \end{align} 
2190: for the ``full S-brane" model, and
2191: \begin{align}
2192: \delta S_{\mathrm{half}} = \lambda \int \dd s e^{X^0(s)} \ , \label{halfi}
2193: \end{align}
2194: for the ``half S-brane" model. Here $X^0$ denotes the target-space time coordinate and the integration is taken over the boundary of the world-sheet parametrized by $s$.
2195: 
2196: There are several different ways to obtain the boundary states. Originally Sen \cite{Sen:2002nu} proposed to obtain the boundary states for \eqref{fulli} by starting with the (compactified) space-like model (boundary sine-Gordon model) and performing the Wick rotation. In the ``half S-brane model", Gutperle and Strominger \cite{Gutperle:2003xf} proposed to use the Wick rotation of the Liouville theory in the zero linear dilaton limit (time-like Liouville theory). 
2197: 
2198: In the coordinate space, the behavior of the boundary states from different prescription shows a different behavior (mainly in the region $ X^0 < 0$), but in the momentum (energy) space, they are related with each other in the zero mode sector. To see this, let us expand the rolling tachyon boundary states as
2199: \begin{align}
2200: |B\rangle &= i\int_C \dd t  \rho (t) |0\rangle  + \sigma (t) \alpha^0_{-1}\bar{\alpha}^0_{-1}|0\rangle + \cdots \cr
2201:  &= i\int \dd\omega \tilde{\rho} (\omega) |\omega \rangle  + \tilde{\sigma} (\omega) \alpha^0_{-1}\bar{\alpha}^0_{-1}|\omega\rangle + \cdots \ .
2202: \end{align}
2203: Since we are dealing with the time-dependent theory based on the analytic continuation, the contour choice will affect the physics. The zero mode density $\rho(\omega)$ has been computed as 
2204: \begin{align}
2205: i \int_{C_{real}}\dd t \rho_{full} (t) e^{i\omega t} &= \left( e^{-i\omega \log\hat{\lambda}} - e^{i\omega \log \hat{\lambda}}\right) \frac{\pi}{\sinh\pi \omega} \cr
2206: i \int_{C_{real}} \dd t \rho_{half} (t) e^{i\omega t} & =  e^{-i\omega \log\hat{\lambda}} \frac{\pi}{\sinh\pi \omega} \cr
2207: i \int_{C_{HH}} \dd t \rho_{full} (t) e^{i\omega t} & =  e^{-i\omega \log\hat{\lambda}} \frac{\pi}{\sinh\pi \omega} \ .
2208: \end{align}
2209: Note that the Hartle-Hawking contour $C_{HH}$ integral of the full S-brane solution coincides with the real contour $C_{real}$ integral of the half S-brane solution (see figure \ref{fig:realHH}). This is intuitively expected because the half S-brane solution describes the later half dynamics of the rolling D-brane (decaying brane) and the Hartle-Hawking contour effectively sets the initial condition at $t=0$ to give a decaying D-brane. We also note that the boundary time-like Liouville field approach directly gives the same zero mode boundary wavefunction for the half S-brane solution. Thus we can conclude that various approaches yield essentially the identical results for the zero mode boundary wavefunction (i.e. coupling to the scalar tachyon mode).
2210: 
2211: \begin{figure}[htbp]
2212:    \begin{center}
2213:     \includegraphics[width=0.5\linewidth,keepaspectratio,clip]{realHH.eps}
2214:     \end{center}
2215:     \caption{Different contour integration gives different boundary states.}
2216:     \label{fig:realHH}
2217: \end{figure}
2218: 
2219: 
2220: 
2221: 
2222: Nevertheless, there are differences in the nonzero mode sectors between the boundary sine-Gordon approach and the boundary time-like Liouville approach. The origin of the difference is that the descendants for the boundary sine-Gordon model is based on the $SU(2)$ current algebra (at the self dual radius) and those for the boundary Liouville theory is based on the Virasoro algebra. For on-shell amplitudes (and energy-momentum tensor) we can gauge away these differences as we will do in section \ref{sec:5-2} to compute the closed string emission rate. However, at least in the two-dimensional noncritical string example, it has been stressed in \cite{Sen:2004yv} that such off-shell boundary states will be important to generate infinitely many conserved charges in addition to the energy momentum tensor. We will revisit the problem later in the discussion of the rolling D-brane, so we will not delve into the details any further at this point and concentrate on the physics associated with the zero mode.
2223: 
2224: 
2225: \subsection{Radiation from rolling tachyon boundary states}\label{sec:5-2}
2226: In this subsection, we would like to study the closed string emission rate from the rolling tachyon by using the exact boundary states. We will present rather a technical aspects of the computation for two reasons. One is that we are going to compare the results of the rolling tachyon and rolling D-brane in later sections in detail. The other is to understand the nontrivial relation between the unitarity (optical theorem) and the open - closed duality, which we revisit in the more nontrivial rolling D-brane case in section \ref{sec:8-2}.
2227: 
2228: Before entering into the computation, we summarize the main physics involved.
2229: \begin{itemize}
2230: 	\item At the one-loop level computation, all the energy of the D0-brane is converted into closed string radiation: the radiation rate shows a power-like divergence.
2231: 	\item Most of the energy is converted into highly massive strings whose mass is effectively cut off by $M\sim 1/g_s$.
2232: 	\item The emitted strings are highly non-relativistic.
2233: 	\item If one considers the D$p$-brane as we increase $p$, the divergence becomes milder, but the spectral density is still power-like and the higher moment diverges.
2234: 	\item The inclusion of the {\it space-like} linear dilaton makes the divergence disappear due to the exponential suppression for the growth of density of states.
2235: 	\item On the contrary, the {\it time-like} linear dilaton (along the rolling tachyon direction) does not affect the divergence. This suggests a first hint of the universality of the decay of unstable D-branes.
2236: \end{itemize}
2237: 
2238: Now we will begin our study on the closed string emission rate from the unstable D-brane. For a slight generalization of section \ref{sec:5-1}, we consider the unstable D-brane in the linear dilaton background. For the boundary states, we will use the one obtained from the time-like Liouville theory because with a time-like linear dilaton, the corresponding boundary states from the boundary sine-Gordon theory is unavailable. For zero time-like linear dilaton limit, however, the two computation agrees as expected.
2239: 
2240:  The dilaton gradient is set
2241: by:
2242: %
2243: \begin{align} \Phi = {1 \over \sqrt{\alpha'}} (Q\, X^0 + {\bf V} \cdot {\bf
2244: X}), \qquad \mbox{where} \qquad Q \equiv \beta - {1 \over \beta} ~
2245: \qquad(\beta \ge 1)~. \label{dilaton}\end{align}
2246: %
2247: This puts the critical dimension $D$ for the bosonic string theory
2248: to be
2249: %
2250: \begin{align} 26 = D - 6 Q^2 + 6 {\bf V}^2, \qquad \mbox{so} \qquad c_{\rm
2251: eff} = 6 Q_\beta^2 - 6 {\bf V}^2~, \end{align}
2252: %
2253: where $Q_\beta \equiv (\beta + 1/\beta)$. The effective central
2254: charge $c_{\rm eff}$ sets the growth of density of closed string
2255: states \cite{Kutasov:1990ua}:
2256: \begin{align}
2257: \rho^{(c)}(M) \sim e^{4\pi \sqrt{\frac{c_{\msc{eff}}}{24}
2258: \alpha' M^2}}
2259: %= e^{4\pi \sqrt{(1-\frac{1}{4} Q^2)\cdot {1 \over 2} \alpha' M^2}}
2260: \label{dos}
2261: \end{align}
2262: %
2263: up to subleading pre-exponential factor of $M$. It grows slower than
2264: the density of states for flat space-time (obtainable by setting
2265: $Q={\bf V} = 0$).
2266: 
2267: 
2268: \subsubsection{closed string emission}\label{sec:5-2-1}
2269: Let us consider the decay of an unstable D-brane in the linear dilaton
2270: background.
2271: % is described by timelike boundary Liouville theory for
2272: %the string time coordinate $X^0$ and spacelike bulk linear dilaton
2273: %theory for the string spatial coordinate.
2274: The radiative transition of a D$p$-brane to a single closed string
2275: state of mass $M$ (set by the integer-valued oscillator level
2276: $N=\widetilde{N}$), whose on-shell energy-momentum $(\omega, {\bf
2277: k})$ is given by
2278: %
2279: \begin{align} \Big(\omega_E - {i Q \over \sqrt{\alpha'}}\Big)^2 - \Big({\bf
2280: k}_E + {i {\bf V} \over \sqrt{\alpha'}} \Big)^2 =(\omega^2 -{\bf
2281: k}^2) =  M^2 \quad \mbox{where} \quad \frac{1}{4}\, \alpha' M^2 = N
2282: - {c_{\rm eff} \over 24}, \end{align}
2283: %
2284: where $(\omega_E, {\bf k}_{E})$ and $(\omega, {\bf k})$ are
2285: energy-momenta in the Einstein and the string frame, respectively.
2286: In string loop perturbation theory, the transition amplitude is
2287: computed by the disk one-point function $\langle \exp ((-i \omega +
2288: \frac{Q}{\sqrt{\al'}}) X^0) \, \exp ((i {\bf k} + \frac{{\bf
2289: V}}{\sqrt{\al'}}) \cdot {\bf X}) \rangle_{\msc{disk}}$ with the
2290: D$p$-brane boundary condition,\footnote{We only consider the case
2291: when the D-brane has Neumann boundary condition in the space-like
2292: linear dilaton direction.} where the vertex operator is separated
2293: into temporal and spatial parts as indicated. The two parts are
2294: factorized in the gauge that no oscillator in temporal direction is
2295: allowed. Consequently, the transition probability ${\cal P}(\omega)$
2296: of the radiative process is governed entirely by the temporal part
2297: (see (3.29) in \cite{Karczmarek:2003xm}):
2298: %
2299: \begin{align} {\cal P}(\omega) = \Big| \left\langle e^{ (- i \omega
2300: +\frac{Q}{\sqrt{\al'}}) X^0} e^{(i {\bf k} + \frac{{\bf
2301: V}}{\sqrt{\al'}})\cdot{\bf X}} \right\rangle_{\msc{disk}} \Big|^2
2302: &= \Big| {1 \over \beta} \Gamma(1 + i \omega\sqrt{\alpha'} \beta)
2303: \Gamma(- i \omega \sqrt{\alpha'} /\beta)\Big|^2 \nn
2304: %
2305: &={\pi^2/\beta^2 \over \sinh (\pi \omega\sqrt{\alpha'}  \beta)
2306: \sinh (\pi \omega\sqrt{\alpha'}  / \beta)}~. \end{align}
2307: %
2308: Then, at leading order in string perturbation theory, the total
2309: number of emitted closed strings from the decay of a D$p$-brane
2310: ($p\ge 1$) extended along ${\bf V}$-direction is computed as
2311: \begin{align} \overline{\cal N} = N_p^2 V_p\sum_M \sqrt{\rho^{(c)}(M)}
2312: \int_{-\infty}^\infty \!{\rmd^{D-1-p} {\bf k} \over (2 \pi)^{D-1-p}}
2313: \, {1 \over 2 \omega} {\cal P}(\omega)~, \end{align} \label{closed}
2314: \noindent where the overall coefficient abbreviates $N_p =
2315: \pi^{\frac{D-4}{4}} (2 \pi)^{\frac{D-2}{4}-p}$ and $V_p$ is the
2316: D$p$-brane volume. In (\ref{closed}), the sum is over all final
2317: closed string states of mass $M$ and of oscillator excitations
2318: symmetric between left- and right-moving sectors. Such oscillator
2319: excitations are equivalent in combinatorics to open string
2320: excitation, so the density of the final states is given by
2321: square-root of \eqref{dos}.
2322: 
2323: Attributed to the Hagedorn growth of the density of states
2324: $\rho^{(c)}(M)$, the total emission number $\overline{\cal N}$ in
2325: (\ref{closed}) (or higher spectral moment) is ultraviolet convergent
2326: so long as linear dilaton has a nonzero spatial component, ${\bf V}
2327: \ne 0$, first observed in \cite{Karczmarek:2003xm}. Notice also that temporal
2328: component of the linear dilaton does not alter the ultraviolet
2329: behavior. This is most readily seen for small ${\bf V}$ by expanding
2330: the density of states. To study anatomy of the ultraviolet behavior,
2331: we shall now perform Fourier transformation and re-express
2332: $\overline{\cal N}$ in the open string channel.
2333: 
2334: 
2335: \subsubsection{open string channel viewpoint}\label{sec:5-2-2}
2336: 
2337: Physical observables such as $\overline{\cal N}$ ought to be
2338: well-defined under the Fourier transform from the closed string
2339: channel to the open string one because
2340: \begin{enumerate}
2341:  \item We start with defining expression of $\overline{\cal N}$,
2342:  consistent with the optical theorem in the closed string channel.
2343: %%%
2344:  \item The expression is closed in the Euclidean signature.
2345: Hence we are free from any subtlety that may arise from analytic
2346: continuations between Euclidean and Lorentzian signature of the
2347: space-time.
2348: \end{enumerate}
2349: %
2350: As in \cite{Karczmarek:2003xm}, we expand the transition probability ${\cal
2351: P}(\omega)$ in convergent power series, whose terms can be
2352: interpreted as D-instantons arrayed along imaginary time
2353: coordinate:
2354: %
2355: \begin{align} {\cal P}(\omega) = {4 \pi^2\over \beta^2} \sum_{n,m=0}^\infty
2356: e^{-2 \pi \alpha' \omega W(m,n)} \end{align}
2357: %
2358: where the location of the D-instantons is denoted as
2359: %
2360: \begin{align} \alpha'  W(m,n) = \sqrt{\alpha'}\Big[\Big(n+{1 \over 2} \Big)
2361: \beta + \Big(m+{1 \over 2} \Big){1 \over \beta}\Big] \ge
2362: \sqrt{\alpha'}~. \end{align}
2363: %
2364: Thus, we take
2365: %
2366: \begin{align} \overline{\cal N} = \Big({2 \pi N_p\over \beta}\Big)^2 V_p
2367: \sum_M \int_{-\infty}^\infty {\rmd^{D-1-p} {\bf k} \over
2368: (2\pi)^{D-1-p}} \sum_{m,n=0}^\infty {1 \over 2 \omega} e^{ - 2 \pi
2369: \alpha' \omega W(m,n)} \end{align}
2370: %
2371: and rewrite each D-instanton contribution parametrically via the
2372: closed string channel modulus $t_c$ as
2373: %using
2374: %
2375: %$ \dsp \int_{-\infty}^{\infty} {\rmd k_0 \over 2 \pi} \,
2376: %\frac{e^{i \xi k_0}}
2377: %{k_0^2+{\bf k}^2+M^2} =
2378: %\frac{1}{2 \om}e^{- \xi \om} $
2379: %;
2380: %
2381: \begin{align} {1 \over 2 \omega} e^{ - 2 \pi \alpha' \omega W(m,n)} &=
2382: %\int_{-\infty}^\infty {\rmd k_0 \over 2 \pi} \, {1 \over k_0^2 +
2383: %\omega^2} e^{2 \pi i k_o n} \nn
2384: %
2385: %&=&
2386: \frac{\pi \alpha'}{2} \int_{-\infty}^\infty {\rmd k_0 \over 2 \pi}
2387: \int_0^\infty \rmd {t_c} \, e^{-2\pi t_c \cdot
2388: \frac{1}{4}\alpha'(k_0^2 + {\bf k}^2 + M^2)} e^{2 \pi i \alpha' k_0
2389: W(m,n)} \ , \end{align}
2390: %
2391: which gives
2392: %
2393: \begin{align} \overline{\cal N} \!\!\! = \cr & \!\!\!\! \Big({2 \pi N_p\over
2394: \beta}\Big)^2 V_p \frac{\pi \alpha'}{2} \sum_{m,n=0}^\infty
2395: \int_0^\infty \rmd {t_c} \int_{-\infty}^\infty {\rmd k_0 \over 2
2396: \pi} \int_{-\infty}^\infty {\rmd^{D-1-p} {\bf k} \over (2
2397: \pi)^{D-1-p}} \, e^{-2\pi t_c \cdot \frac{1}{4}\alpha'(k_0^2 + {\bf
2398: k}^2)} e^{2 \pi i \alpha' k_0 W(m,n)}\nn
2399: %
2400: & \hskip3.7cm \times \sum_M \sqrt{\rho^{(c)}(M)} e^{-2\pi t_c \cdot
2401: \frac{1}{4}\alpha' M^2}. \end{align}
2402: %
2403: Here, we exchanged order of summations and integrations, and first
2404: performed integrals over off-shell momenta $(k_0, {\bf k})$ and sum
2405: over mass level $M$. The sum over $M$ yields modular covariant
2406: partition function $Z^{(c)}(q_c)$ in terms of the Dedekind eta
2407: function:
2408: %
2409: \begin{align} Z^{(c)} (q_c) &\equiv \sum_M \sqrt{\rho^{(c)}(M)} \, \,
2410: q_c^{\frac{1}{4}\alpha'M^2} \qquad \mbox{where} \qquad q_c \equiv
2411: e^{-2\pi t_c} \nn &= \eta^{-(D-2)}(q_c)~. \end{align}
2412: %
2413: Integrations over the $(D-p)$-dimensional momenta $(k_0, {\bf k})$ yield
2414: $(2 \pi^4 \alpha' t_c)^{-(D-p)/2}$ times Gaussian damping factor
2415: $e^{-2\pi \alpha' W^2(m,n)/ t_c}$. We now perform modular
2416: transformation to the open string channel $t_c = 1/t_o$, where $t_o$
2417: is modulus of the open string channel and $q_o \equiv e^{- 2\pi
2418: t_o}$. Putting all these together, we finally have
2419: %
2420: \begin{align} \overline{\cal N} =  C_p \, V_p
2421:  \sum_{m,n=0}^\infty \int_0^\infty {\rmd t_o \over t_o} t_o^{-p/2}\, e^{- 2\pi
2422: t_o \alpha' W^2(m,n) } \, \eta^{-(D-2)} (q_o), \label{openexp}\end{align}
2423: %
2424: with $C_p = \Big({2 \pi N_p\over \beta}\Big)^2 \frac{\pi
2425: \alpha'}{2}(2\alpha'\pi^4)^{-\frac{D-p}{2}}$, reproducing the result
2426: reported in \cite{Karczmarek:2003xm}. As it stands, the final expression
2427: \eqref{openexp} is at odd to the intuition based on, for example, the
2428: Schwinger pair production in (time-dependent) electric field, since
2429: the integral over the open string modulus $t_o$ is still intact. If
2430: the total emission number can be interpreted as arising from on-shell
2431: two-particle branch cut in the open string channel, the modulus
2432: integral ought to be absent! Therefore, To understand underlying
2433: physics better, we shall now compute the cylinder amplitude directly
2434: and then extract the imaginary part via the optical theorem.
2435: 
2436: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2437: \subsubsection{Lorentzian cylinder amplitude}\label{sec:5-2-3}
2438: Unitarity and optical theorem thereof, combined with the open-closed
2439: string channel duality, should enable us to extract the emission
2440: number $\overline{\cal N}$ of closed strings from decaying D$p$-brane
2441: as the imaginary part of the cylinder amplitude. In the closed
2442: string channel diagram, the computation reduces to \eqref{closed}, as
2443: in quantum field theory. It is, however, somewhat nontrivial to
2444: evaluate the imaginary part of the cylinder amplitude directly from
2445: the open string channel. Here we present the {\sl ab initio}
2446: derivation, refining that in the text of \cite{Karczmarek:2003xm}, by starting
2447: with manifestly well-defined Lorentzian cylinder amplitude.
2448: 
2449: We begin with the cylinder amplitude in the closed string channel in
2450: which both the world-sheet and the target space-time signatures are
2451: taken Lorentzian. Omitting overall numerical factors for the moment,
2452: the amplitude is given by
2453: \begin{align}
2454: Z_{\rm cylinder} = i \pi \alpha' V_p \int_{s_c^{\rm IR}}^{s_c^{\rm
2455: UV}} \rmd s_c \int_{-\infty}^{\infty} \frac{\rmd \omega_L}{2\pi} \,
2456: \frac{\pi^2/\beta^2 \cdot
2457: q_c^{-(1-i\hat{\epsilon})^2\frac{1}{4}\alpha'\omega_L^2}
2458: }{\sinh(\pi\beta\omega_L \sqrt{\alpha'}) \sinh(\pi\omega_L
2459: \sqrt{\alpha'}/\beta)} \, Z_{\cM}^{(c)}(q_c) \ , \label{cl}
2460: \end{align}
2461: where $q_c = e^{2\pi i \tau_c}$ with $\tau_c = s_c + i\epsilon$, and
2462: $Z_{\cM}(q_c)$ represents the contribution from the closed string
2463: zero-modes and oscillator parts.\footnote{We are using different
2464: normalization for modulus parameters from \cite{Karczmarek:2003xm}: $t$(KLMS) =
2465: $(\pi/4)t$(here). In addition, they adopted $\alpha'=1$
2466: convention.} The Lorentzian world-sheet is regularized by $i
2467: \epsilon$ prescription, while the Lorentzian space-time is
2468: regularized by $-i \hat{\epsilon}$-prescription. $s_c^{\rm {UV}}$
2469: ($s_c^{{\rm IR}}$) is an ultraviolet (infrared) regulator of the closed
2470: string channel modulus. With these prescriptions, the integral over
2471: $\om_L$ is convergent so long as $2 \hat{\ep} s_c^{\rm UV} > \ep >0$
2472: is retained.
2473: 
2474: 
2475: Defining the open string modular parameter
2476: as $q_o = e^{-2\pi i \tau_o}$ where
2477: $\tau_o = s_o - i \epsilon$ with $s_o = 1/s_c$, one can rewrite
2478: \eqref{cl} in terms of open string channel energy $\omega_L'$ as
2479: \begin{align}
2480: Z_{\rm cylinder} &=  V_p \int_{s_o^{\rm UV}}^{s_o^{\rm IR}} {\rmd
2481: s_o} \int_{-\infty}^{\infty} \rmd \omega'_L \left(i \pi \alpha'
2482: \int_{-\infty}^{\infty} \rmd \omega_L \frac{\cos(\pi \alpha'
2483: \omega_L \omega'_L)}{\sinh(\pi\beta \omega_L \sqrt{\alpha'} )
2484: \sinh(\pi\omega_L\sqrt{\alpha'}/\beta)} \right) \nonumber \\
2485: & \hskip4cm \times \,
2486: q_o^{-(1+i\hat{\epsilon}')^2\frac{1}{4}\alpha'{\omega_L'}^2}
2487: Z_{\cM}^{(o)}(q_o) \, ,
2488: \end{align}
2489: %%%
2490: where $s^{\rm IR}_o\equiv 1/s^{\rm UV}_c$, $s^{\rm UV}_o\equiv
2491: 1/s^{\rm IR}_c$ are the cut-off's in the open string modulus.
2492: %%%
2493: As opposed to the closed string channel, we have to adopt the
2494: $+i\hat{\ep}'$-prescription for the Lorentzian space-time, and
2495: the above integral is well-defined as long as
2496: $2 \hat{\ep}' s^{\rm UV}_o > \ep$.
2497: %%%
2498: The expression in the large parenthesis yields the open string
2499: density of states, $\rho^{(o)}(\omega'_L)$. It is infrared divergent
2500: at $\omega_L= 0$. To regularize it, we subtract minimally the double
2501: pole\footnote{This subtraction does not affect the imaginary part
2502: of the partition function we are primarily interested in.} so that
2503: \begin{align}
2504: \rho^{(o)}(\omega'_L)_{\mathrm{reg}} &= i \pi \alpha'
2505: \int_{-\infty}^{\infty} \rmd \omega_L \left(\frac{\cos(\pi\alpha'
2506: \omega_L\omega'_L)}{\sinh(\pi\beta\omega_L \sqrt{\alpha'})
2507: \sinh(\pi\omega_L \sqrt{\alpha'}/\beta)} - \frac{1}{\pi^2
2508: \alpha'\omega_L^2} \right) \cr &= -{2}\partial_{\omega'_L} \log
2509: S_\beta\left({Q}_\beta + i\sqrt{\alpha'}\omega'_L\right) \ , \label{dens}
2510: \end{align}
2511: where the `$q$-Gamma function' $S_\beta(x)$ is defined by\footnote
2512:   {Here the normalization of variable $x$ differs with factor 2
2513:    from the one given in \cite{FZZ}.}
2514: \begin{align}
2515: -\partial_x \log S_\beta (x) = \int_{-\infty}^{\infty} \rmd t
2516: \left(\frac{\cosh((x-Q_\beta)t)}{2\sinh(\beta t)\sinh(t/\beta)}
2517: -\frac{1}{2t^2} \right)\
2518: \end{align}
2519: for $\mathrm{Re}(x) < 2 Q_\beta$ and analytically continued to the
2520: whole complex plane.\footnote{Notice that the Lorentzian density
2521: \eqref{dens} is well-defined without the analytic continuation. We
2522: stress that this should be contrasted against the approach of
2523: \cite{Karczmarek:2003xm}.} See, for example, \cite{FZZ,Nakayama:2004vk}.
2524: 
2525: Now we perform the Wick-rotation both in the target space and on the
2526: world-sheet. First, Wick rotate the open string channel energy as
2527: $\om'_L\, \rightarrow\, e^{i(\frac{\pi}{2}-0)} \om'_L$ and set
2528: $\om'_L = i \om'$ $(\om' \in \br)$. Then, we can safely Wick rotate
2529: the world-sheet Schwinger parameter as $s_o \, \rightarrow \, - i
2530: t_o$ ($t>0$). Notice that we will need to perform the Euclidean
2531: rotation in opposite direction for the closed and the open string
2532: channels due to the difference of the $i\ep$-prescription.
2533: %first in the target space
2534: %as $\omega'_L \to e^{i(\frac{\pi}{2} -0)} \omega'_L = i\omega'$,
2535: %and then, on the worldsheet as $s_o \to - it_o$ $(t_o>0)$.
2536: %Notice that we have to make the space-time Wick-rotation
2537: %in opposite directions for the closed and open string channels
2538: %due to the difference of $i\ep$-prescription.
2539:  There is no obstruction in such contour deformation because
2540:  $\partial_x \log S_\beta (x) $ has poles only on the real axis.
2541:  We will see that this is specific to the decaying D-brane situation
2542:  and do not hold generally. In fact, in section \ref{sec:8-2}
2543:  dealing with the rolling D-branes,
2544:  we shall show that there exist extra contributions from crossing poles
2545:  in the course of the contour rotation and that their contributions are
2546:  essential for maintaining the unitarity.
2547:  After Wick rotating the world-sheet,
2548: the cylinder amplitude in the open string sector
2549:  is given by
2550: \begin{align}
2551: Z_{\rm cylinder} = -2 V_p \int_{0}^\infty \rmd t_o
2552: \int_{(1-i0)\mathbb{R}} \rmd \omega' \partial_{\omega'} \log S_\beta
2553: \left({Q_\beta} -\sqrt{\alpha'}
2554: \omega'\right)q_o^{\frac{1}{4}\alpha'{\omega'}^2} Z_M^{(o)}(q_o) \ .
2555: \end{align}
2556: 
2557: Imaginary part of the partition function comes from the simple poles
2558: of the $q$-Gamma function $S_\beta \left(Q_\beta -\sqrt{\alpha'}
2559: \omega'\right)$ at $\frac{1}{2}\omega' = W(m,n)$
2560: %(m+\frac{1}{2})\beta + (n+\frac{1}{2})\beta^{-1}$
2561: for $n,m \in \mathbb{Z}_{\ge 0} $ and simple zeros for $n,m \in
2562: \mathbb{Z}_{< 0}$. Therefore, collecting imaginary parts from the
2563: contour integration over $\omega'$ and applying the optical theorem,
2564: we finally obtain
2565: \begin{align}
2566: \overline{\cal N} =  \mathrm{Im}\, Z_{\rm cylinder} = C_p \, V_p
2567: \sum_{n,m=0}^\infty \int_0^\infty {\rmd t_o \over t_o}
2568: t_o^{-\frac{p}{2}}\, e^{-2\pi t_o \alpha' W^2 (m,n)} \,
2569: %((n+1/2)\beta + (m+1/2)/\beta)^2 }
2570:  \eta^{-(D-2)}(q_o) \ ,
2571: \end{align}
2572: where we have evaluated the free oscillator part explicitly and
2573: reinstated overall numerical factors.
2574: %As was demonstrated in \cite{Karczmarek:2003xm}
2575: This is in perfect agreement with \eqref{dilaton}, and it may be
2576: interpreted as a nontrivial check of unitarity and open-closed
2577: duality in the Lorentzian signature.
2578: 
2579: 
2580: 
2581: \subsubsection{D-brane decay in two-dimensional string theory}\label{sec:5-2-4}
2582: In a similar method, one can compute the spectral observables from
2583: the D-brane decay in two-dimensional string theory \cite{Klebanov:2003km}. The boundary
2584: state for the unstable D-brane in two-dimension is given by the
2585: ZZ-brane boundary state \cite{Zamolodchikov:2001ah}:
2586: \begin{align}
2587: \langle e^{(i k + 2/\sqrt{\alpha'})\phi} \rangle_{\msc{disk}} =
2588: \mu^{-\frac{i}{2} \sqrt{\alpha'} k} \frac{2
2589: \sqrt{\pi}}{\Gamma(1-ik\sqrt{\alpha'})\Gamma(ik\sqrt{\alpha'})} \ .
2590: \end{align}
2591: Combining it with the rolling tachyon boundary states, the total
2592: emission number of closed string is given by
2593: \begin{align}
2594: \overline{\cal N} = N_o^2 \int^{\infty}_{0} \rmd k \int^{\infty}_{0}
2595: \frac{\rmd \omega}{2\omega}  {\cal P} (\omega, k) \delta(\omega-k) \
2596: ,
2597: \end{align}
2598: where the on-shell condition $\omega = k$ is imposed, and the
2599: transition probability is
2600: \begin{align}
2601: {\cal P} (\omega, k)  = \left| \langle e^{- i\omega X^{0}} e^{
2602: (ik+2/\sqrt{\alpha'})\phi} \rangle_{\msc{disk}} \right|^2 =
2603: \frac{\sinh^2(\pi k\sqrt{\alpha'})}{\sinh^2(\pi \omega
2604: \sqrt{\alpha'})} \ .
2605: \end{align}
2606: We see that, after performing the $k$-integration, the resultant
2607: total emission number is ultraviolet divergent.
2608: 
2609: To express $\overline{\cal N}$ in open string channel, we repeat the
2610: analysis of section \ref{sec:5-2-3} and expand the transition probability in
2611: arrays of imaginary D-instantons. The result is
2612: \begin{align}
2613: \overline{\cal N} &= N_o^2 \sum_{m,n = 0}^\infty \int_0^\infty \rmd
2614: k \int_0^\infty \rmd t_c \int_{-\infty}^{\infty} \frac{\rmd
2615: k_0}{2\pi} e^{-2\pi t_c \cdot \frac{1}{4}\alpha' (k_0^2 + k^2)}
2616: e^{2\pi i \alpha' k_0 W(m,n)} \sinh^2(\pi k\sqrt{\alpha'}) \Big\vert_{\beta \rightarrow 1} \cr
2617: %
2618: &= N_o^2 \sum_{m,n = 0}^\infty \int_{0}^\infty \frac{\rmd t_o}{t_o}
2619: \Big({1 \over q_o} - 1 \Big) q_o^{ \alpha' W^2(m,n)}
2620: \Big\vert_{\beta \rightarrow 1}
2621:  \ , \label{twodim}
2622: \end{align}
2623: where we have reinstated $W(m,n)$ for the purpose of
2624: regularization.\footnote{Because of the subtraction of singular
2625: vector in $(1/q_o - 1)$, the resultant amplitude is {\it
2626: non-unitary}.} The expression exhibits ultraviolet divergence as
2627: $t_o \to \infty$.
2628: 
2629: On the other hand, it is possible to obtain the same radiation rate
2630: from the direct evaluation of the imaginary part of the Lorentzian
2631: cylinder amplitude in the open-string channel as was done in section
2632:  \ref{sec:5-2-3}:
2633: \begin{align}
2634: Z_{\rm cylinder} = i N_o^2 \int_0^\infty \rmd s_c
2635: \int_{-\infty}^{\infty} \frac{\rmd \omega_L}{2\pi} \int_0^{\infty}
2636: \frac{\rmd k}{2\pi} \, \frac{\sinh(\pi \sqrt{\alpha'} k)^2
2637: }{\sinh(\pi \sqrt{\alpha'} \omega_L)^2}\,
2638: q_c^{\frac{1}{4}\alpha'(-\omega^2_L + k^2)}  \ .
2639: \end{align}
2640: After rewriting the open string density by the $q$-Gamma function as
2641: in section \ref{sec:5-2-3}, we obtain open string channel expression of the
2642: partition function. We then find the imaginary part from the poles
2643: located at $\frac{1}{2}\omega' = W(m,n)$, and reproduce
2644: \eqref{twodim}. This confirms that the partition function is
2645: manifestly unitary, obeying the optical theorem. Here again, the
2646: regularization $\beta \to 1$ is implicit.
2647: 
2648: \subsubsection{ZZ brane decay in various dimensions}\label{sec:5-2-5}
2649: It is possible to generalize the discussion of section \ref{sec:5-2-4} for ZZ branes in various dimensions by introducing the time-like linear dilaton theory. If we write the dilaton slope of the Liouville and time-like linear dilaton direction as $V_{\phi} = b+{b}^{-1}$ and $V_{t} = \beta - \beta^{-1}$ respectively, the criticality condition is given by
2650: \begin{align}
2651: 26 = D+6V_{\phi}^2 - V_{t}^2 \ . \label{crirr}
2652: \end{align}
2653: We can combine the one-point function for the ZZ brane (with general $b$) and the decaying D-brane boundary states for the boundary time-like Liouville theory to compute the radiation rate as was studied in \cite{He:2006bm}.
2654: A similar cancellation as we discussed in section \ref{sec:5-2-1} gives the UV power-like structure of the closed string radiation rate.\footnote{Technically speaking, for $D>26$, we encounter a closed string IR divergence \cite{He:2006bm}.} The result suggests again the universality of the decaying D-brane spectrum.
2655: 
2656: In this construction, owing to the criticality condition \eqref{crirr}, the existence of the time-like dilaton is unavoidable. In the following, we study the decay of the ZZ brane in $\mathcal{N}=2$ Liouville theory (or D0-brane Euclidean $SL(2;\br)/U(1)$ coset model. See section \ref{sec:6} for more details) to study more realistic models, where $\beta = 1$ (i.e. flat limit) is feasible (see \cite{Israel:2006ip} for a particular case. This subsection is based on a generalization of their results).
2657: 
2658: The absolute square of the boundary wavefunction for the $\mathcal{N}=2$ ZZ-brane is given by
2659: \begin{align}
2660: |\Psi(p,m)|^2 = \delta_{m,\bar{m}} \frac{\sinh(2\pi p)\sinh(2\pi p/k)}{\cosh(2\pi p) + \cos(\pi m)} \ , \label{zzwv}
2661: \end{align}
2662: while that for the ($\mathcal{N}=1$ supersymmetric) rolling D-brane with time-like linear dilaton $V_{t} = \beta - \beta^{-1}$ is 
2663: \begin{align}
2664: |\Psi(E)^2| = \frac{1}{\sinh(\beta E) \sinh(\beta^{-1}E)} \ . \label{rrd}
2665: \end{align}
2666: We can evaluate the on-shell ($E^2 = M^2 + \frac{p^2}{2k} + \frac{m^2}{2k}$) emission with fixed transverse mass $M$ as 
2667: \begin{align}
2668: N(M) &= \int \dd p \sum_m |\Psi(p,m)^2||\Psi(E(p,m))|^2 \cr
2669:      &\sim \int \dd p e^{\frac{\pi}{k}p - \pi(\beta + \beta^{-1})\sqrt{M^2+\frac{p^2}{2k}}} \cr
2670: 	& \sim e^{-2\pi M \sqrt{(\frac{\beta+\beta^{-1}}{2})^2 - \frac{1}{2k}}} \ ,
2671: \end{align}
2672: where in the last line we have used the saddle point approximation. 
2673: 
2674: On the other hand, the density of states for the emitted closed string for large $M$ is given by
2675: \begin{align}
2676: \sqrt{\rho^{(c)}} \sim e^{4\pi M \sqrt{\frac{c_{\mathrm{eff}}}{24}}} = e^{2\pi M \sqrt{(\frac{\beta+\beta^{-1}}{2})^2 - \frac{1}{2k}}} \ .
2677: \end{align}
2678: Thus, we see an exact cancellation of the exponential part of the closed emission rate, leaving us with a familiar power-like universal closed string emission rate.
2679: 
2680: We have several comments here
2681: \begin{itemize}
2682: 	\item We can analyse the bosonic case in the same way. The first difference is $k$ in \eqref{zzwv} should be replaced with $\kappa-2$. The second difference is $E$ in \eqref{rrd} should be replaced with $\sqrt{2}E$. The final closed string emission rate changes, as a consequence, to $N(M) \sim e^{-2\pi \sqrt{(\beta+\beta^{-1})^2/2 - \frac{1}{2(\kappa-2)}}}$, which will cancel against the bosonic Hagedorn density of states $\sqrt{\rho^{(c)}} \sim e^{4\pi M \sqrt{\frac{c_{\mathrm{eff}}}{24}}} \sim e^{2\pi \sqrt{(\beta+\beta^{-1})^2/2 - \frac{1}{2(\kappa-2)}}}$.
2683: 	\item For simplicity, we studied the emission rate from the closed string perspective. The open string computation like we did in section \ref{sec:5-2-2}, \ref{sec:5-2-3} is straightforward, and we will not repeat it here.
2684: 	\item The conclusion here is independent of the level $k$ of the $SL(2;\br)/(1)$, which, on one hand, suggests a universality of the D-brane decay. On the other hand, it seems curious to observe that nothing special happens at $k=1$, where we expect a ``black hole - string transition". As we will see in section \ref{sec:7}, \ref{sec:8} in detail, the rolling (or Euclidean hairpin) D-brane captures or probes the ``black hole - string transition". We will return to this question in section \ref{sec:9}.
2685: \end{itemize} 
2686: \subsubsection{electric field and long string formation}\label{sec:5-2-6}
2687: One simple generalization of the rolling D-brane was, as we studied in section 5.1.3, the inclusion of the linear dilaton. Another simple generalization is to introduce constant electric field on the D-brane, i.e. we introduce the fundamental string charge \cite{Mukhopadhyay:2002en,Rey:2003xs}.
2688: 
2689: In order to introduce the constant electric field (say $F^{01} = \epsilon$) on the D-brane, we can use the stringy version of the Lorentz boost. The successive applications of T-duality, Lorentz boost and inverse T-duality, we end up with the boundary states with electric flux. Operationally, the transformation is
2690: \begin{align}
2691: |0 \rangle \to \gamma |0\rangle \ , \ \  t\to \gamma^{-1} t \ , \ \ \omega \to \gamma \omega  \cr
2692: \begin{pmatrix} \alpha^0 \\ \alpha^1 \end{pmatrix} \to \Lambda^{-1} \begin{pmatrix} \alpha^0 \\ \alpha^1 \end{pmatrix} \ , \ \ \ \begin{pmatrix} \bar{\alpha}^0 \\ \bar{\alpha}^1 \end{pmatrix} \to \Lambda \begin{pmatrix} \bar{\alpha}^0 \\ \bar{\alpha}^1 \end{pmatrix} \ , 
2693: \end{align}
2694: where 
2695: \begin{align}
2696: \Lambda = \gamma \begin{pmatrix}  1& \epsilon \\ \epsilon & 1 \end{pmatrix} \ , \ \ \gamma = \frac{1}{\sqrt{1-\epsilon^2}} \ .
2697: \end{align}
2698: 
2699: From this transformation law, the energy momentum tensor can be easily read as
2700: \begin{align}
2701: T_{00} & \sim E \gamma \cr
2702: T_{01} & \sim -Ee^2\gamma -E \gamma^{-1} \exp(-\gamma^{-1}t) \cr
2703: T_{11} & \sim -E \gamma^{-1} \exp(-\gamma^{-1}t) \ ,
2704: \end{align}
2705: in the $t\to \infty$ limit. The study of the closed string radiation from the boundary states is straightforward. When $x^1$ direction is noncompact, the result is 
2706: \begin{align}
2707: \langle N \rangle = \sum_M \int \dd k \frac{|\Psi(\omega_{k,M})|^2}{2\omega_{k,M}} \simeq \int^{\infty} \dd M \sqrt{\rho^{(c)}(M)}e^{-2\pi \gamma M} = \int^{\infty}\dd M e^{-2\pi (\gamma-1) M}
2708: \end{align}
2709: and the total emission rate is exponentially suppressed essentially due to the Lorentz time delay \cite{Nagami:2003yz,Nagami:2003mr}.
2710: 
2711: Now let us suppose that $x^1$ direction is compactified with the radius $R$. In this case, we have to sum over the winding mode:
2712: \begin{align}
2713: N(M) = \sum_w \int \dd k \frac{|\Psi(\omega_{k,M})|^2}{2\omega_{k,M}} \simeq \sum_w\int \dd k e^{- 2\pi\gamma (\sqrt{(wR)^2 + k^2 + M^2}-\epsilon Rw)} \ .
2714: \end{align}
2715: For large $M$, the summation over $w$ can be evaluated by the saddle point methods, which leads to 
2716: \begin{align}
2717: \langle N \rangle 
2718: \sim \int^\infty \dd M \sqrt{\rho^{(c)}(M)} N(M) \sim \int^\infty \dd M M^{\beta}
2719: \end{align}
2720: We recover the power-like behavior of the emission rate \cite{Gutperle:2004be}.\footnote{The power dependence $\beta$ is determined from the details of the model e.g. dimensionality of the D-brane and the details of the internal CFT etc.} The computation reveals that the winding mode dominates the emission rate in the electrified D-brane decay. Mathematically, this is due to the (T-dualized) Lorentz invariance in the $R\to 0$ limit. Physically, the decay of the D-brane produces many long macroscopic strings as a final decay product, which has a cosmological significance as we will review in section \ref{sec:5-5}.
2721: 
2722: \subsection{Classical correspondence}\label{sec:5-3}
2723: The Dirac-Born-Infeld form of the rolling tachyon effective action \eqref{tDBI} suggests a possible geometrical interpretation of the open string tachyon condensation. Such a geometrical interpretation of the rolling tachyon process would shed a new light upon our understanding of the nature of the open string tachyon and its condensation. It would also provide a guiding principle for a geometrical interpretation of the closed string tachyon condensation, for qualitative properties of the closed string tachyon condensation are poorly understood compared with the open string tachyon condensation.
2724: 
2725: In \cite{Kutasov:2004dj}, an interesting connection between the D-brane motion in the (near horizon) NS5-brane background and the rolling tachyon dynamics was pointed out. Since the NS5-brane has a tension proportional to $1/g_s^2$, in perturbative string theories, we can regard it as a fixed background, in which the D-brane, whose tension is proportional to $1/g_s$ moves. In other words, in the perturbative string theories, the probe D-brane approximation is good and trustful.
2726: 
2727: The effective action for the D-brane motion in NSNS-background (i.e. without any R-R fields), is given by the Dirac-Born-Infeld action
2728: \begin{align}
2729: S = -T_p\int \dd^{p+1} \sigma e^{-\Phi}\sqrt{-\det(X^*[G+B]_{\mu\nu})} \ ,
2730: \end{align}
2731: where $X^*[G+B]$ denotes the pullback to the D$p$-brane world-volume.
2732: 
2733: As proposed in \cite{Kutasov:2004dj}, let us consider the D0-brane motion in near horizon NS5-brane geometry \eqref{NH ext NS5}.\footnote{If one considers a homogeneous motion of the D-brane, the net result does not depend on the spacial dimension of the D-brane. We also assume that D-brane sits at a point in the internal space $\mathbb{S}^3$.} Let us fix the world-sheet reparametrization invariance by taking the static gauge $\sigma^0 = t$. In this gauge, the DBI action reduces to
2734: \begin{align}
2735: S = -T_0 \int \dd t e^{\frac{\rho}{\sqrt{2k}}}\sqrt{1-\dot{\rho}^2} \ , \label{DBIr}
2736: \end{align}
2737: where dot denotes the derivative with respect to $\sigma_0 = t$, and we have rescaled the radial direction $\rho$ so that we have a canonical kinetic term.
2738: 
2739: Let us compare the effective action for the radion field $\rho$ \eqref{DBIr} with the open string tachyon effective action \eqref{tDBI}. It is almost clear in the large (negative) region of $\rho$, these two expressions essentially coincide with each other.\footnote{If one take $k=2$, the coincidence becomes exact including the numerical factor in the tachyon (radion) potential.} This is the classical ``tachyon - radion correspondence": one can identify the effective action for the rolling tachyon problem with the effective action for the rolling D-brane in the NS5-brane, or linear dilaton, background. The ``radion field" $\rho$ plays the role of the tachyon field $T$ here.
2740: Note, however, that the radion field is actually not tachyonic, although it has run-away potential, nor has an unstable extremum in the potential because it is a massless field at the tree level. 
2741: 
2742: One can readily solve the classical equation of motion based on the action \eqref{DBIr} as
2743: \begin{align}
2744: e^{-\frac{\rho}{\sqrt{2k}}} = c \cosh\left(\frac{t}{\sqrt{2k}}\right) \ , \label{seoms}
2745: \end{align}
2746: which agrees with the late time behavior of the rolling tachyon problem \eqref{seom}. The energy momentum tensor can be read as 
2747: \begin{align}
2748: T_{00} &= E\delta(\rho-\rho_0(t)) \cr
2749: T_{0\rho} &= E \tanh\left(\frac{t}{\sqrt{2k}}\right) \delta(\rho-\rho_0(t)) \cr
2750: T_{ij} &= -E \mathrm{sech^2}\left(\frac{t}{\sqrt{2k}}\right) \delta(\rho-\rho_0(t)) \delta_{ij} \ \ \ (i,j = 1,\cdots,p) \ ,
2751: \end{align}
2752: where $\rho_0(t)$ is the classical solution of the radion motion \eqref{seoms}. As expected, the energy momentum tends to that for a pressure-less dust as $t\to \infty$. The $(0\rho)$ component has a natural interpretation as the momentum transfer in the $\rho$ direction because the decaying D-brane moves in the $\rho$ direction almost at the speed of light as $t\to \infty$.
2753: 
2754: What is the end point of the ``radion condensation"? In the case of the open string tachyon condensation, Sen's conjecture states that we end up with the closed string vacuum, where the open string excitation becomes infinitely massive and disappear from the physical spectrum. From the effective field theory approach taken here, it is difficult to establish this statement in a satisfactory manner because in the large $\rho$ regime, the effective string coupling becomes larger due to the linear dilaton gradient. One way to study this might be to uplift the system to M-theory (e.g. by using the interpolating metric proposed in \cite{Aharony:1998ub}). The subsequent physics, however, is intuitively clear: the D-brane will be absorbed into the NS5-brane and form a non-threshold bound states. The open string spectrum on the D-brane should be modified so that it matches with the excitation on the bound states.\footnote{It would be an interesting open problem to study the tachyon - radion correspondence from the open string field theory and prove the analogue of Sen's conjecture.}
2755: 
2756: 
2757: There are several generalizations of the problem. One interesting question is whether we can obtain the effective DBI action having the exactly identical potential with the rolling tachyon not only in the large $\rho$ region. This is possible by considering an array of the NS5-brane on $\br^3\times \mathbb{S}^1$ rather than the stack of NS5-branes in $\br^4$ \cite{Kutasov:2004ct}. Because of the oppositely-directed attractive force between two NS5-branes, the potential of the D-brane can have a local extremum:
2758: \begin{align}
2759: S = -T_0 \int \dd t \frac{1}{\cosh\frac{\rho}{\sqrt{2k}}}\sqrt{1-\dot{\rho}^2} \ ,\label{aDBIr}
2760: \end{align}
2761: which completely agrees with \eqref{tDBI}.
2762: Unfortunately, unlike the NS5-branes on $\br^4$, the exact quantization of the rolling D-brane in this geometry is unavailable.\footnote{The exact boundary states for {\it static} (unstable) brane in a similar background has been constructed in \cite{Eguchi:2004ik}, which reproduces the mass of the geometrical tachyon (i.e. radion).}
2763: 
2764: Another interesting generalization is to consider the D-brane motion in the non-extremal black NS5-brane background. Interestingly, after a simple coordinate transformation, the classical motion of D-brane in the non-extremal NS5-brane (outside of the horizon) is identical to that in the extremal NS5-brane.
2765: To see this we note that, by introducing `tachyon' variable $Y \equiv \log \sinh \rho$,
2766: DBI Lagrangian of the D0-brane can be cast to that of rolling
2767: tachyon:
2768: %
2769: \begin{align} L_{\rm D0} &= - e^{-\Phi} \sqrt{\left({\dd s \over \dd
2770: t}\right)^2} = - V(Y) \sqrt{1 - \dot{Y}^2} \qquad \mbox{where}
2771: \qquad
2772:  V(Y) = M_0 \, e^Y ~, \label{nonextremal D0} \end{align}
2773: if we restricted ourselves to the region outside of the horizon.
2774: An important point is that,
2775: in sharp contrast to the extremal background \eqref{NH ext NS5}, the
2776: dilaton is finite everywhere. Thus, the strong coupling singularity
2777: is now capped off by the horizon. The construction of the exact boundary states for the rolling D-brane in the two-dimensional black hole (or non-extremal NS5-brane) is one of the main themes of this thesis.
2778: 
2779: For another example of exactly solvable deformation, one can introduce constant electric fields as we did in the rolling tachyon example. This has been studied in \cite{Nakayama:2004ge}, where we have constructed exact boundary states and have shown the correspondence between the electrified rolling tachyon problem and the electrified rolling radion problem even with $\alpha' \sim 1/k$ corrections. As yet another generalization, the rotating D-brane solution in NS5-brane background has been also studied in \cite{Kutasov:2004dj}, which could be regarded as a rotational Lorentz boosted solution as pointed out in \cite{Nakayama:2004ge}, but the exact boundary state is yet to be constructed. Other classical studies of D-brane motion in related background include \cite{Yavartanoo:2004wb,Panigrahi:2004qr,Ghodsi:2004wn,Sahakyan:2004cq,Toumbas:2004fe,Bak:2004tp,Chen:2004vw,Kluson:2005qx,Lapan:2005qz,Thomas:2005am,Thomas:2005fw,Kluson:2005dr,Kluson:2005eb,Kluson:2005zw,Papantonopoulos:2006eg,Okuyama:2006zr,Gumjudpai:2006hg}.
2780: 
2781: Before concluding this subsection, we would like to stress again that the correspondence at the level of the effective action is only valid in the large $\rho$ or $T$ regime, where the effective action analysis loses its validity because the effective string coupling grows there. Therefore, the quantum correspondence we will prove in later sections, based on the one-loop string perturbation theory, is actually not so obvious, and we should rather regard it as a highly nontrivial statement of the universality of the properties of decaying D-branes.
2782: 
2783: \subsection{Quantum correspondence}\label{sec:5-4}
2784: 
2785: So far, we have mainly discussed the classical correspondence between the rolling tachyon problem and the rolling radion problem at the level of the effective action. Aside from the debate over the effectiveness of the rolling tachyon DBI-like action \eqref{tDBI} , we have one tunable parameter $k$ in the rolling radion problem, so it is important to analyse a possible $k$ dependence of this correspondence. 
2786: 
2787: We know that $1/k$ measures the $\alpha'$ corrections to the background geometry from the discussion in section \ref{sec:2}. When $k$ becomes larger, the classical geometry, and hence, the DBI action is more trustful. On the other hand, when $k$ becomes smaller, the geometry shows large $\alpha'$ corrections and the effective action approach may break down. Especially, the exact correspondence at the level of the effective action requires $k=2$, which is rather in a strongly coupled regime.\footnote{In the bosonic case, we need to set $k=1$.} 
2788: 
2789: In particular, if one considers the two-dimensional black hole geometry (as the non-extremal NS5-branes background), the appearance of the stretched horizon blurs the geometry. In addition, we expect a ``black hole - string transition" at $k=1$. It is of utmost interest to probe such a phase transition from the rolling D-brane. 
2790: 
2791: In the following sections, we will construct the exact boundary states for such rolling D-branes in NS5-brane background, and reveal the nature of the $1/k \sim \alpha'$ corrections to the tachyon - radion correspondence. After the construction of the exact boundary states, we study the closed string radiation rate as we did in the rolling tachyon case in section \ref{sec:5-2} and compare the results. 
2792: 
2793: For convenience, we summarize our main physical results here \cite{Nakayama:2005pk}:
2794: 
2795: \begin{enumerate}
2796: 	\item The closed string emission rate from the rolling D-brane (which will be computed in section \ref{sec:8-2}) yields exactly the same behavior as that from the rolling tachyon (which was computed in section \ref{sec:5-2}). Especially, the power-like behavior of the spectrum density does not depend on $k$ (up to an overall normalization). This is true as long as $k>1$, and confirms the tachyon - radion corresponding from the exact boundary states.
2797: 	\item Independence of the extra parameter $k$, which even governs the world-sheet (stringy) $\alpha'$ correction suggests a universal nature of the decaying D-brane: all the energy of the D-brane will be radiated as a gas of closed strings, whose dominant contribution comes from the highly massive (long) strings. If one introduces the fundamental string charge, as an electric flux, the dominant contributions from the rolling D-brane again comes from the winding strings as we have seen in the rolling tachyon problem in section \ref{sec:5-2-6}.
2798: 	\item The situation changes drastically if one studies the case $k<1$. The closed string emission rate is exponentially suppressed, and the tachyon - radion correspondence breaks down. This is in accord with the ``black hole - string transition" at $k=1$ discussed in section \ref{sec:4}. Our result is the first physical manifestation of the ``black hole - string transition" in the two-dimensional black hole probed by the rolling D-brane. 
2799: \end{enumerate}
2800: \subsection{Cosmological implications}\label{sec:5-5}
2801: 
2802: From the early days of its invention, the rolling tachyon system has also been studied in the context of the cosmological applications. In particular, the realization of the inflation in string theory has attracted more and more attention recently with increasing evidence for the existence of such period in the history of our universe (see \cite{Linde:2005dd} and references therein). Indeed, one of the simplest proposals for the inflation from the string theory is the tachyon inflation, where the (open string) tachyon plays the role of the inflaton \cite{Kofman:2002rh,Frolov:2002rr,Li:2002et,Fairbairn:2002yp,Shiu:2002qe}. The tachyon - radion correspondence discussed so far enables us to consider varieties of radion (or geometrical tachyon) inflation. From the classical tachyon - radion correspondence, many features of the tachyon inflation can be translated into that of the radion inflation with more generalities \cite{Thomas:2005fu}.
2803: 
2804: The starting point of the tachyon (radion) inflation is (minimal) coupling of the DBI-like action \eqref{tDBI} \eqref{DBIr} to the gravity:
2805: \begin{align}
2806: L_{\mathrm{eff}} = \sqrt{-g} \left(\frac{R}{16\pi G} - V(T)\sqrt{1+g^{\mu\nu}\partial_\mu T \partial_\nu T} \right) \ . \label{frwl}
2807: \end{align}
2808: For our realistic application, we consider the four-dimensional (non-compact) space-time, and $8\pi G = M_p^{-2}$ with the four-dimensional Planck constant $M_p$. Under the assumption of the Friedman-Robertson-Walker isotropic universe, the four dimensional metric can be written as\footnote{Since the inflation flattens the space in an exponential manner, we have assumed a flat space universe for simplicity.}
2809: \begin{align}
2810: \dd s^2 = -\dd t^2 + a(t)^2 \dd x^2_i  \ \ \ (i = 1,2,3) \ .
2811: \end{align}
2812: 
2813: We begin with the equation of motion for $T$:
2814: \begin{align}
2815: \frac{\ddot{T}}{1-\dot{T}^2} + 3H\dot{T} + \frac{V'}{V} = 0 \ ,
2816: \end{align}
2817: where the prime denotes the derivative with respect to $T$ (i.e. $V' = \partial V(T)/\partial T$) and the dot denotes the time derivative. $H$ here denotes the Hubble parameter $H \equiv \dot{a}/{a}$.
2818: In the slow-roll approximation (i.e. $\dot{T}\gg 1$), the Friedman equation reads
2819: \begin{align}
2820: H^2 = \left(\frac{\dot{a}}{a}\right)^2 = \frac{V(T)}{3M_p^2} \ ,
2821: \end{align}
2822: and the slow-roll equation reduces to
2823: \begin{align}
2824: 3H\dot{T} = \frac{V'(T)}{V} \ .
2825: \end{align}
2826: For the slow-roll parameter $\eta \sim (H')^2/H^4$ to be small enough, we must require
2827: \begin{align}
2828: H^2 \gg \frac{(V')^2}{V^2} \ .
2829: \end{align}
2830: 
2831: Suppose our (geometrical) tachyon potential has a local extremum as is the case with the rolling tachyon and the geometrical tachyon in the array of NS5-brane backgrounds. Inflation near the local extremum is possible if $H^2 \gg |m^2|$, where $m^2$ is the mass for the (geometric) tachyon. The condition is equivalent to
2832: \begin{align}
2833:  \frac{g_s}{v l_s} \gg \frac{C}{k l_s} \ , \label{gom}
2834: \end{align}
2835: where $v$ is the volume of the compactification,\footnote{We are assuming a direct product type compactification. If we consider the warped compactification, we can relax the condition.} and if we kept track of every numerical factor, we could find $C\sim 260$. In the original rolling tachyon problem, $k = 2$ and it is difficult to find a consistent solution in the perturbative string theory while maintaining the COBE normalization $H/M_p \sim 10^{-5}$ \cite{Kofman:2002rh,Frolov:2002rr}. In the geometric tachyon, we have one parameter $k$, and if we choose large enough $k$, it is possible to satisfy the condition \eqref{gom} consistent with the COBE normalization. We can also satisfy the slow-roll condition in the geometric tachyon. For instance, $\eta \ll 1$ is equivalent to the condition $H \gg |m|$ for $T < O(1)$. The key point here is that we have an extra tunable parameter $k$ to obtain a sustainable inflation in the case of the geometric tachyon unlike the original rolling tachyon cosmology, where such a tunable parameter is absent.
2836: 
2837: Nevertheless, we still have a serious drawback of this rolling tachyon (radion) type cosmology as pointed out in \cite{Kofman:2002rh,Frolov:2002rr,Shiu:2002qe}. The problem is related to how the inflation will end. Since the effective potential for the rolling tachyon (radion) runs away exponentially as $T \to \infty$, there is no minimum for the tachyon to oscillate. Therefore, it is allegedly impossible to reheat the universe to produce various matters, i.e. after the tachyon inflation we end up with an empty universe, which is of course unacceptable.
2838: 
2839: Again in the contex of the geometric tachyon (radion), we can avoid this reheating problem by preparing a ring of the NS5-branes and evolution of the D-brane inside the ring \cite{Thomas:2005fu}. The effective potential has global minima and the oscillation around the minima produces the reheating needed to produce matter and hence our galaxies. 
2840: 
2841: We can continue this line of reasoning and study for instance the spectral index of the cosmic microwave background etc, but this is not the main scope of this thesis. Rather, we would like to point out how inaccurate this kind of effective action analysis for the rolling D-brane  is {\it after coupling to the massive closed string sector}. Especially, the stringy treatment of the decay of the D-brane completely changes the nature of the reheating from such rolling tachyon (or D-brane) systems.
2842: 
2843: Let us begin with the following illustrative toy example. In the above discussions, couplings of the (decaying or rolling) D-brane to higher massive stringy modes (except for graviton) have been neglected. In the usual field theory, we expect corrections to the effective action of order $\sim M_{p}^2/M^2$, where $M^2$ denotes the mass of the fields integrated out. The point is that we have to sum over infinitely many massive fields: e.g. in the Kaluza-Klein theory, $M^2 \propto n^2$, where $n$ denotes the internal momenta, so the summation over $n$ schematically gives 
2844: \begin{align}
2845: \sum_n \frac{1}{M_n^2} \sim \sum_n \frac{1}{n^2} = \frac{\pi^2}{6} \ ,
2846: \end{align}
2847: which is finite. However, in string theory, the number of massive string modes grows exponentially $\rho(M) \sim \exp(\beta_{\rm Hg} M)$ as we discussed the Hagedorn temperature in section \ref{sec:4-1}. Thus the summation over all the massive modes with the coupling $\sim 1/M_s^2$ clearly diverges.
2848: 
2849: In reality, the coupling to the massive closed string sectors is much softer and the exponentially suppressed as $\exp(-\beta M)$. The case-by-case computation is needed to see  which exponential factor governs, but from our results (summarized in section \ref{sec:5-5}) it seems universal that the exponential part cancels out and the closed string backreaction is characterized by a power-like behavior irrespective of the superficial strength of the $\alpha' \sim 1/k$ corrections.\footnote{It is also interesting to note that the exponential suppression of the higher massive modes occurs when $k<1$ in the regime where the supergravity approximation is invalid.}
2850: 
2851: In this way, we can conclude that the reheating of the universe through the rolling tachyon - radion is rather effective than one might expect from the naively truncated effective action. As the direct calculation shows, almost all of the energy is radiated as the closed strings without any need for the oscillation around the extremum.\footnote{As we have discussed in section, \ref{sec:5-2}, the emission rate is power-like finite for higher dimensional branes. However, this does not mean an effectiveness of the truncated effective action (DBI+FRW such as \eqref{frwl}) to discuss the closed string backreaction. It  just means that it is more effective to decay by disconnecting patches of D-brane as D0 particles (assuming it is uncharged). Mathematically, it is just an artefact of the one-point decay and the one-point decay is no more effective than the higher-point decay.} The actual problem, therefore, is how we can transmit energy from the radiated (massive) closed string to the standard model sector. This problem is rather model dependent and we will not study it any further in detail here (see e.g. \cite{Kofman:2005yz,Frey:2005jk,Chialva:2005zy,Chen:2006ni} for recent studies).
2852: 
2853: As we have studied in section \ref{sec:5-2}, the final decay product of the rolling tachyon and rolling D-brane is highly probable to be long closed strings. This can be directly seen by assigning fundamental string charges to the unstable D-branes, but without assigning such charges, intuitively it is expected to be so by considering a pair production. This is in good agreement with the  usual Kibble mechanism of producing long macroscopic strings in the universe: causally disconnected region creates long strings and they evolve independently. It would be of great interest to study this problem quantitatively from the string theory viewpoint and determine a remaining density of cosmic strings associated with the D-brane decay (see \cite{Polchinski:2004ia} for a review of cosmic strings from the superstring theory). Such studies will verify or even exclude the geometric tachyon inflation. It is also of great importance to revisit the reheating process of various D-brane inflation scenarios to see whether the classical oscillatory contribution is really dominant over the emission of the highly massive string modes.
2854: 
2855: \newpage
2856: \sectiono{D-branes in Two-dimensional Black Hole}\label{sec:6}
2857: We now begin with our studies on D-branes in the two-dimensional black hole background. In this section, we review the D-branes in the Euclidean two-dimensional black hole. The organization of the section is as follows. In section \ref{sec:6-1}, we classically analyze the D-branes in the two-dimensional black hole and derive the mini-superspace boundary wavefunction. In section \ref{sec:6-2}, we review the exact boundary states describing the D-branes in the two-dimensional black hole system. 
2858: \subsection{Classical D-branes}\label{sec:6-1}
2859: \subsubsection{DBI analysis}\label{sec:6-1-1}
2860: The classification of the D-brane in general curved backgrounds is given by the solution of the Dirac-Born-Infeld action coupled with Chern-Simons action.\footnote{Of course, one could imagine unstable D-branes whose effective action is {\it not} given by the DBI action + Chern-Simons, but they are outside the scope of our discussion.} The total effective action is
2861: \begin{align}
2862: S = -\mu_p \int \dd^{p+1}\xi e^{-\Phi}\sqrt{-\det(G_{ab}+B_{ab}+F_{ab})} +i\mu_p\int e^{F_2+B_2} \wedge \sum_q C_q \ , \label{DBICS}
2863: \end{align}
2864: where the summation over $C_q$ should be taken over all R-R fields in the theory we are considering. 
2865: 
2866: In this section, we study the D-branes in the Euclidean two-dimensional black hole:
2867: \begin{align}
2868: \dd s^2 =  k\al' (\tanh^2\rho \, \dd \theta^2 + \dd
2869: \rho^2) \ , 
2870: \qquad e^{2\Phi} = \frac{k}{\mu \cosh^2 \rho} \ .
2871: \end{align}
2872: In the Euclidean two-dimensional black hole background,
2873:  there exist no Kalb-Ramond $B_{\mu\nu}$ field nor the R-R fields, so the effective action is simply given by the DBI term with possible electro-magnetic flux $F_{\mu\nu}$ on it. Since we are in the Euclidean signature, the DBI action \eqref{DBICS} should be Wick-rotated in an appropriate manner:\footnote{Our ``Wick rotation" here is nothing but adding a (dummy) extra decoupling time direction and set trivial Neumann boundary condition along the time. In particular, we will assume $F + B$ is real.}
2874: \begin{align}
2875: S^E= \mu_p \int \dd^{p}\xi e^{-\Phi}\sqrt{\det(G_{ab}+B_{ab}+F_{ab})} \ .
2876: \end{align}
2877: 
2878: We begin with the (Euclidean) D0-brane. The D0-brane is a point particle and the DBI action on it is simply given by
2879: \begin{align}
2880: S_{\mathrm{D}0} = \mu_0 e^{-\Phi} \propto \cosh\rho \ .
2881: \end{align}
2882: It is clear that the extremum of the action is obtained when $\rho=0$. Thus we conclude that the D0-brane is localized at the tip of the cigar.
2883: 
2884: Next we study the D1-brane. The DBI action for the D1-brane is given by
2885: \begin{align}
2886: S_{\mathrm{D}1} = \mu_1 \int \dd\theta \cosh\rho(\theta) \sqrt{\rho'(\theta)^2 + \tanh^2\rho(\theta)} \ ,
2887: \end{align}
2888: where we have fixed the reparametrization invariance by using the gauge $\xi = \theta$. The equation of motion is easily solved from the ``energy" conservation:
2889: \begin{align}
2890: \text{const} = \cosh\rho \frac{\tanh^2{\rho}}{\sqrt{(\rho')^2 + \tanh^2\rho}} \ 
2891: \end{align}
2892: as
2893: \begin{align}
2894: \sinh(\rho) \cos(\theta -\theta_0) = \sinh\rho_0 \ . \label{trsce}
2895: \end{align}
2896: 
2897: For later purposes, we note that if one uses the complex coordinate
2898: \begin{align}
2899:  u = \sinh\rho e^{i\theta} \ , \ \ \bar{u} = \sinh\rho e^{-i\theta} \ ,
2900: \end{align}
2901: the classical trajectory \eqref{trsce} takes the form of a straight line in the complex  $u$ plane. This can be also seen from the fact that the DBI action takes the flat form
2902: \begin{align}
2903: S_{D1} = \mu_1 \int \dd\xi \sqrt{\frac{du}{d\xi}\frac{d\bar{u}}{d\xi}} \ 
2904: \end{align}
2905: in this coordinate.
2906: 
2907: We finally examine the D2-brane. In this case, we can introduce a magnetic flux $F_{\rho\theta} = f(\rho,\theta)$. By fixing the reparametrization invariance as $\xi_1 = \rho$, $\xi_2 =\theta$, the DBI action reads
2908: \begin{align}
2909: S_{\mathrm{D}2} = \mu_2 \int d\theta d\rho \cosh\rho \sqrt{\tanh^2\rho + f^2(\rho,\theta)} \ .
2910: \end{align}
2911: From the Gauss law constraint, we have
2912: \begin{align}
2913:  c = \frac{\cosh\rho f(\rho,\theta)}{\sqrt{\tanh^2\rho + f^2(\rho,\theta)}} \ ,
2914: \end{align}
2915: which determines the magnetic flux as 
2916: \begin{align}
2917: f^2(\rho,\theta) = \frac{c^2\tanh^2\rho}{c^2 - \cosh^2\rho} \ . \label{magfl}
2918: \end{align}
2919: If $c>1$, the D2-brane partially wraps the cigar and has a boundary at $\rho = \mathrm{arccosh}(c)$ because at that value of $\rho$, the magnetic field blows up. On the other hand, if $c<1$, the D2-brane wraps the whole cigar. In the latter case, the magnetic field on the D2-brane induces a D0-brane charge near the tip of the cigar, which should be quantized. Writing $c = \sin\sigma \le 1$, we obtain the classical quantization condition as 
2920: \begin{align}
2921: \frac{\sigma - \sigma'}{2\pi} k \in \bz \ .
2922: \end{align}
2923: 
2924: In quite a similar fashion, we can also study the classical D-branes in the T-dualized trumpet background:
2925: \begin{align}
2926: \dd s^2 = \dd\rho^2 + \frac{1}{\tanh^2\rho} \dd\tilde{\theta}^2 \ , \ \ e^{\Phi} = \frac{k}{\mu \sinh \rho} \ .
2927: \end{align}
2928:  Since the discussion is completely in parallel, we only present the results. 
2929:  
2930:  The D0-brane (probably D1-brane?) could be localized at $\rho = 0$. Since $\rho=0$ is a singularity in the trumpet geometry, the presence of such D-branes are not obvious at all. Formally, we can regard it as a T-dual of the D0-brane of the cigar geometry.
2931:  
2932:  The D1-brane is given by the solution of the DBI action
2933:  \begin{align}
2934: S_{\mathrm{D}1} = \mu_{\mathrm{D}1} \int \dd\tilde{\theta}\sinh\rho\sqrt{\frac{1}{\tanh^2\rho}+({\rho}')^2} \ ,
2935:  \end{align}
2936:  in the static gauge.
2937: The solution is given by 
2938: \begin{align}
2939: \cosh\rho \cos(\tilde{\theta}-\tilde{\theta}_0) = \gamma \ .
2940: \end{align}
2941: when $\gamma>1$, the D1-brane is connected, while when $\gamma<1$, the D1-branes go through the singularity and possibly they become disconnected. Naturally, the D1-brane in the trumpet geometry is regarded as a T-dual of the D2-brane in the cigar geometry. The parameter $\gamma$ corresponds to the parameter $c$ in the cigar geometry.\footnote{The parameter $\tilde{\theta}_0$ can be T-dualized to the holonomy of the gauge field $A_0$ in the cigar. Since the D2-brane has a nontrivial fundamental group $\pi_1 = \mathbb{Z}$, different $A_0$ gives a different D-brane (for $c>1$).}
2942: 
2943: The D2-brane is classified by the solution of the DBI action
2944: \begin{align}
2945: S_{D2} = \mu_{D2}\int \dd\rho \dd\tilde{\theta} \sinh\rho \sqrt{\frac{1}{\tanh^2\rho} + F^2} \ .
2946: \end{align}
2947: The Gauss law constraint gives 
2948: \begin{align}
2949:  F^2 = \frac{\beta^2}{\tanh^2\rho(\sinh^2\rho-\beta^2)} \ .
2950: \end{align}
2951: The D2-brane always has a boundary at $\rho = \mathrm{arcsinh}(\beta)$. The D2-brane in the trumpet geometry naturally corresponds to the T-dual of the D1-brane in the cigar. The parameter identification is obviously given by $\beta = \sinh\rho_0$ appearing in \eqref{trsce}.
2952: 
2953: \subsubsection{group theoretical viewpoint}\label{sec:6-1-2}
2954: In section \ref{sec:6-1-1}, we have studied the classical D-brane in the Euclidean two-dimensional black hole from the effective DBI action. Since the two-dimensional black hole system can be realized as the $SL(2;\br)/U(1)$ coset model, we can study the classification of the D-brane from the gauged WZNW model \cite{Maldacena:2001ky,Gawedzki:2001ye,Elitzur:2001qd,Fredenhagen:2001kw,Ishikawa:2001zu,Yogendran:2004dm}. Indeed all the D-branes discussed in section \ref{sec:6-1-1} descend from the branes in the parent $SL(2;\br)$ WZNW model. 
2955: 
2956: The starting point is the D-branes in the parent $SL(2;\br)$ WZNW model. We focus on the maximally symmetric D-branes for technical simplicity. As we proceed, we will see that the maximally symmetric D-branes are enough to obtain all the D-branes constructed from the DBI analysis done in section \ref{sec:6-1-1}. The maximally symmetric D-branes in the WZNW model are classified by the (twined) conjugacy class of the group $G$ with a possible quantization condition \cite{Kato:1996nu,Alekseev:1998mc,Birke:1999ik,Behrend:1999bn,Felder:1999ka}. We call them A-branes (conjugacy class) and B-branes (twined conjugacy class) respectively.
2957: 
2958: In our $SL(2;\br)$ group with the Euler angle parametrization $g = e^{i\sigma_2 \frac{t-\theta}{2}} e^{\rho\sigma_1}e^{i\sigma_2\frac{t+\theta}{2}}$ , the conjugacy class is given by 
2959: \begin{align}
2960: \mathrm{Tr} (g) = 2 \cos{t} \cosh{\rho} \equiv 2\kappa , \label{conj}
2961: \end{align}
2962: and the twined conjugacy class is given by
2963: \begin{align}
2964: \mathrm{Tr} (\sigma_1 g) = 2\cos{\theta} \sinh{\rho} \equiv 2\kappa' \ . \label{tconj}
2965: \end{align}
2966: up to conjugation.
2967: 
2968: The D-branes in the axial coset model is obtained by gauging $g$ by $hgh$, where $h^a= e^{i\sigma_2 a}$ in our case. For A-brane, we have to sum over the gauge orbit of the parent D-brane \label{conj} parametrized by $\kappa$ in order to obtain a gauge invariant object. The gauge transformation of the conjugacy class is given by
2969: \begin{align}
2970: \mathrm{Tr}(h^agh^a) = 2\cos(t+a) \cosh{\rho} ,
2971: \end{align}
2972: so the gauge invariant orbit of the parent D-brane is given by
2973: \begin{align}
2974: \cosh{\rho} \ge \kappa  \ . \label{d2wzw}
2975: \end{align}
2976: Projecting it down on the coset coordinate (in the gauge $t=0$) is now trivial, and we have obtained the D2-brane wrapped (partially) around the cigar whose world volume is restricted by the condition \eqref{d2wzw}. %In a special case $\kappa=0$, we have a D0-brane localized at the tip of the cigar ($\rho=0$). 
2977: The shapes of the A-branes obtained here are in complete agreement with the ones obtained from the DBI analysis in section \ref{sec:6-1-1}. The precise parameter identification is $c=\kappa$ for the D2-brane.\footnote{There are several independent ways to justify this parameter identification. For instance, one can show it directly from  the detailed study on the boundary conditions of the gauged WZNW model with boundaries. In \cite{Walton:2002db}, they have shown that the parameter $\kappa$ is indeed the field strength appearing in the effective action of the D-brane at the boundary by using the T-duality technique. Their study of the $SU(2)/U(1)$ model can be translated to our Euclidean $SL(2;\br)/U(1)$ model with no essential modifications.}  We also note that A-brane is invariant under the isometry of the coset in this construction.
2978: 
2979: Similarly from the parent B-brane, we can construct the D1-brane of the coset. In this case, since the twined conjugacy class is already gauge invariant, we can directly project \eqref{tconj} down onto the coset coordinate. The resulting D1-brane trajectory is given by
2980: \begin{align}
2981:  \sinh{\rho}\cos{\theta} = \kappa' \ ,
2982: \end{align}
2983: which is nothing but the one obtained in \eqref{trsce} from the DBI analysis (with $\theta_0 =0$). The B-brane constructed in this way breaks the isometry of the coset, so it has a Nambu-Goldstone mode along the $\theta$ direction. This corresponds to the rotation of $\sigma_1$ and $\sigma_3$ in the definition of the 
2984:  twined conjugacy class \eqref{tconj}.
2985: 
2986: 
2987: We could repeat the same analysis for the vector coset ($\sim$ trumpet geometry). Since the argument is completely in parallel, we skip the detailed discussion, and simply note that the results agree with the DBI analysis.
2988:  
2989: \subsubsection{mini-superspace boundary wavefunction}\label{sec:6-1-3}
2990: In the context of the string theory, D-branes can be described either from the open string viewpoint or from the closed string viewpoint (i.e. channel duality). Technically, this is achieved by the modular transformation of the cylinder amplitudes. The boundary state $|B\rangle$ is defined by
2991: \begin{align}
2992: Z_{\mathrm{cylinder}} \equiv \int \dd t_o \mathrm{Tr}_o e^{-\pi H_o t_o} = \int \frac{\dd t_c}{t_c} \langle B| e^{-\pi H_ct_c}| B\rangle \ ,
2993: \end{align}
2994: where $H_o = L_0$ is the open string Hamiltonian while $H_c = L_0 + \bar{L}_0$ is the closed string Hamiltonian. The boundary state $|B\rangle$ satisfies the gluing condition 
2995: \begin{align}
2996: (L_{n} -  \bar{L}_{-n}) |B \rangle = 0 , 
2997: \end{align}
2998: for the energy-momentum tensor (and similar gluing conditions for any other conserved currents if any: see section \ref{sec:6-2} for further details).
2999: 
3000: At the level of the minisuperspace approximation, the boundary states can be seen as the coupling of the D-brane to the closed string zero mode:
3001: \begin{align}
3002: \langle B |_{\mathrm{mini}} = \int_0^\infty \frac{\dd p}{2\pi} \Psi_{0}(p,n) \langle\langle p,n| \ , \label{miniwv}
3003: \end{align}
3004: where $|p,n\rangle\rangle$ is the so-called Ishibashi state \cite{Ishibashi:1988kg} associated with the primary states $|p,n\rangle$ (see section \ref{sec:6-2} for details), but in the mini-superspace approximation, there is no difference between the two $|p,n\rangle\rangle \sim |p,n\rangle$ because they are different only in the non-zero mode sector. The semiclassical boundary wavefunction $\Psi_{0} (p,n)$ is obtained from the overwrap between the D-brane and the primary state $|p,n\rangle$ as
3005: \begin{align}
3006: \Psi_0(p,n) = \langle B|p,n\rangle \ . 
3007: \end{align}
3008: The explicit form of the primary state $|p,n\rangle$ and the classical trajectory have been given in the form of the minisuperspace approximation as we have studied in section \ref{sec:3-2-3} and section \ref{sec:6-1-1}.
3009: 
3010: In the following, we compute the minisuperspace boundary wavefunction $\Psi_0$ for each D-branes studied in section \ref{sec:6-1-1}. The results will be compared with the proposed exact boundary states in section \ref{sec:6-2}. We expect that they will agree with each other in the semi-classical limit ($k \to \infty$), and indeed they do as we will see.
3011: 
3012: Let us begin with the D0-brane. Classically, the D0-brane is localized at the tip of the cigar $\rho = 0$, and the boundary wavefunction is simply given by the minisuperspace wavefunction for $|p,n\rangle$ evaluated at $\rho = 0$. From the explicit minisuperspace wavefunction \eqref{ef}, we can easily derive
3013: \begin{align}
3014: \Psi_0^{\mathrm{D}0}(p,n) = - \delta_{n,0} \frac{\Gamma^2(-j)}{\Gamma(-2j-1)} = -\delta_{n,0} \frac{\Gamma^2(\frac{1}{2}-\frac{ip}{2})}{\Gamma(-ip)} \ . \label{minico}
3015: \end{align}
3016: We note that D0-brane does not couple to the momentum mode along $\theta$, which is consistent with the interpretation that the D0-brane is an A-brane (see section \ref{sec:6-1-2}). The exact analysis shows that it couples to the winding mode and the discrete states localized near the tip of the cigar, but the minisuperspace analysis cannot capture them.
3017: 
3018: Next let us consider the D1-brane. Classically, the D1-brane has the shape of the hairpin. 
3019:  A semiclassical D-brane boundary wavefunction is the weighted sum of the
3020: wavefunction of closed string states restricted to the location of
3021: the D-brane. In the mini-superspace approximation, as is implicit in
3022: \cite{Ribault:2003ss}, the weighted sum equals to the overlap between the
3023: mini-superspace wavefunction and the delta function constraint
3024: enforcing $(\rho, \theta)$ coordinates over the hairpin trajectory
3025: \eqref{trsce} (with respect to the volume element \eqref{vol cigar}).
3026: The result is
3027: \begin{align}
3028: & \int_0^\infty \sinh \! \rho \, \dd \sinh \! \rho
3029: \int_{-\frac{\pi}{2}+\theta_0}^{\frac{\pi}{2}+\theta_0} \dd\theta\,
3030: \delta \Big(\cos (\theta-\theta_0) \sinh \rho-\sinh \rho_0 \Big)
3031: \phi^p_{n}(\rho,\theta) \cr &= \int_{-\frac{\pi}{2}}^{\frac{\pi}{2}}
3032: \!\dd \theta' \, \frac{\sinh \rho_0}{\cos^2 \theta'}
3033: \phi^p_{n}(\widehat{\rho}(\rho_0,\theta'),\theta') e^{i n \theta_0}
3034: , 
3035: %\label{overlap phi}
3036: \end{align}
3037: where $\theta'= (\theta-\theta_0)$ and
3038: $\widehat{\rho}(\rho_0,\theta')$ refers to the solution of $\cos
3039: \theta' \sinh \rho= \sinh \rho_0$. Using the decomposition
3040: \eqref{decomp ef}, we are then to evaluate integrals:
3041: \begin{align}
3042:  & \int_{-\frac{\pi}{2}}^{\frac{\pi}{2}} \dd \theta \,
3043: \frac{\sinh \rho_0}{\cos^2 \theta} \,
3044: \phi^p_{L,n}(\hat{\rho}(\rho_0,\theta),\theta) =
3045: \frac{2\pi\Gamma(ip)} {\Gamma\left(\frac{1}{2}+\frac{ip+n}{2}\right)
3046: \Gamma\left(\frac{1}{2}+\frac{ip-n}{2}\right)} \, e^{-ip\rho_0} ~,
3047: \nn
3048: %%%
3049:  & \int_{-\frac{\pi}{2}}^{\frac{\pi}{2}} \dd \theta \,
3050: \frac{\sinh \rho_0}{\cos^2 \theta} \,
3051: \phi^p_{R,n}(\hat{\rho}(\rho_0,\theta),\theta) =
3052: \frac{2\pi\Gamma(-ip)}
3053: {\Gamma\left(\frac{1}{2}-\frac{ip+n}{2}\right)
3054: \Gamma\left(\frac{1}{2}-\frac{ip-n}{2}\right)} \, e^{+ ip\rho_0} ~.
3055: \label{evaluation overlap phi}
3056: \end{align}
3057: Details of the computation are relegated in Appendix \ref{mini}. Using the
3058: mini-superspace reflection amplitude \eqref{cref amp}, we then obtain
3059: \begin{align}
3060: &\Psi^{(0)}_{\rm D1}(\rho_0,\theta_0;p,n) =  \frac{2\pi\Gamma(ip)}
3061: {\Gamma\left(\frac{1}{2}+\frac{ip+n}{2}\right)
3062: \Gamma\left(\frac{1}{2}+\frac{ip-n}{2}\right)} \, e^{in\theta_0}
3063: \left(e^{-ip\rho_0} + (-1)^n e^{+ip\rho_0}\right)~. \label{hairpin
3064: D1 classical}
3065: \end{align}
3066: The D1-brane couples to the momentum mode as is clear from the geometry, which is consistent with the interpretation that the D1-brane is a B-brane (see section \ref{sec:6-1-2}).
3067: 
3068: %We see the result \eqref{hairpin D1 classical} reproduces the exact
3069: %result \eqref{hairpin D1} {\sl modulo the factor}
3070: %$\Gamma\left(1+i\frac{p}{k}\right)$. Importantly, this missing
3071: %factor depends on $k$ (measured in string unit) and hence
3072: %corresponds precisely to the corrections due to string worldsheet
3073: %effects. The mini-superspace approximation sets $k \rightarrow
3074: %\infty$, so this factor is consistently dropped out. Equivalently,
3075: %this missing factor can be reinstated to the D-brane boundary wave
3076: %function by demanding consistency of the wave function in the
3077: %mini-superspace approximation with the exact reflection amplitude
3078: %\eqref{qref amp}.
3079: 
3080: Finally, we study the D2-brane. The D2-brane is parametrized by the parameter $c$ appearing in the amount of the magnetic flux \eqref{magfl}. Since the qualitative features of the D2-brane seem to be different for $c>1$, and $c<1$, it is natural to study them separately. Since the mini-superspace analysis for the D2-brane has not been available in the literature, we would like to present it slightly in detail here.\footnote{The author would like to thank S.~Ribault for stimulating discussions on this problem.}
3081: 
3082: Let us begin with the case when $c>1$. The D2-brane only partially wraps the cigar because at $\rho = \mathrm{arccosh}(c)$, the field strength diverges.
3083: We parametrize $c = \cosh r_0$. 
3084: 
3085: Since the D2-brane couples to the winding states, the minisuperspace analysis is only possible for the zero winding sector.\footnote{We could avoid this problem in the T-dual picture, which will be discussed later.} The (zero momentum/winding) minisuperspace wavefunction is given by 
3086: \begin{align}
3087: \phi_{p,m=0}(\rho) &= -\frac{\Gamma^2(-j)}{\Gamma(-2j-1)}F(j+1,-j;1;-\sinh^2\rho) \cr &= (\sinh\rho)^{-1-ip} F\left(\frac{1}{2} + \frac{ip}{2}, \frac{1}{2} + \frac{ip}{2}; 1+ ip;-\frac{1}{\sinh^2\rho}\right) \cr &+ \frac{\Gamma(ip)\Gamma^2(\frac{1}{2}-\frac{ip}{2})}{\Gamma(-ip)\Gamma^2(\frac{1}{2}+\frac{ip}{2})}(\sinh\rho)^{-1+ip} F\left(\frac{1}{2} - \frac{ip}{2}, \frac{1}{2} - \frac{ip}{2}; 1- ip;-\frac{1}{\sinh^2\rho}\right) \ ,
3088: \end{align}
3089: where $j = -\frac{1}{2} + \frac{ip}{2}$.
3090: 
3091: The boundary wavefunction in the minisuperspace approximation is given by
3092: \begin{align}
3093: \Psi_{2}(r_0)^{\mathrm{mini}} = \int_{r_0}^{\infty} \dd\rho \cosh\rho \frac{\sinh\rho}{\sqrt{\cosh^2\rho-\cosh^2r_0}} \phi_p(\rho) \ .
3094: \end{align}
3095: Now we can perform the integration as follows
3096: \begin{align}
3097: & \int_{r_0}^\infty \dd\rho \cosh\rho  \frac{\sinh\rho}{\sqrt{\cosh^2\rho-\cosh^2r_0}} (\sinh\rho)^{-1-ip} F\left(\frac{1}{2}+\frac{ip}{2},\frac{1}{2}+\frac{ip}{2};1+ip;-\frac{1}{\sinh^2\rho}\right) \cr
3098: &= \frac{\Gamma(ip+1)}{\Gamma(\frac{1}{2}+\frac{ip}{2})^2} \sum_{n=0}^{\infty} (-1)^n (\sinh^2 r_0)^{-n-\frac{ip}{2}}\frac{\sqrt{\pi}}{2}\frac{\Gamma(n+\frac{ip}{2})\Gamma(\frac{1}{2}+\frac{ip}{2}+n)}{\Gamma(ip+1+n)n!} \cr
3099: &= \frac{\Gamma(\frac{ip}{2})}{\Gamma(\frac{1}{2}+\frac{ip}{2})} (\sinh^2 r_0)^{-\frac{ip}{2}} \frac{\sqrt{\pi}}{2} F\left(\frac{ip}{2},\frac{1}{2}+\frac{ip}{2};ip+1,-\frac{1}{\sinh^2 r_0}\right) \cr
3100: &= \frac{\pi\Gamma(ip)}{\Gamma(\frac{1}{2}+\frac{ip}{2})^2} e^{-ipr_0} \ . 
3101: \end{align}
3102: We refer to the appendix \ref{mini} for the last equality (see also \cite{Nakayama:2005pk}).
3103: Combining it with the second integration that can be treated in the same manner, we obtain 
3104: \begin{align}
3105: \Psi_2(r_0)^{\mathrm{mini}} = \frac{\pi\Gamma(ip)}{\Gamma(\frac{1}{2}+\frac{ip}{2})^2} \cos(pr_0) \ . \label{direce}
3106: \end{align}
3107: We can see that the boundary wavefunction for the partially wrapped D2-brane is consistent with the class 2 boundary wavefunction proposed in \cite{Fotopoulos:2004ut} at least for the zero winding sector (see section \ref{sec:6-2} for details).
3108: 
3109: We can repeat our analysis when $\beta \le 1$ and reproduces the minisuperspace limit of the class 3 boundary states in the zero-winding sector.
3110: In the T-dual picture, (partially wrapped) D2-brane in the cigar geometry is supposed to be given by the D1-brane in the trumpet geometry.\footnote{One subtle point of the trumpet geometry is that the semiclassical limit is unclear. We regard $k\to \infty$ as the semiclassical limit for $\rho$ direction, but $\tilde{\theta}$ direction is apparently not. We will neglect this subtlety for a moment.}
3111: 
3112: Let us now move on to the minisuperspace wavefunction for the D1-brane in the trumpet geometry.
3113:  From the semiclassical DBI action
3114: \begin{align}
3115: L = \sinh\rho \sqrt{\dot{\rho}^2 + k^{-2} \coth^2\rho} \ ,
3116: \end{align}
3117: the equation of motion is easily integrated with the help of the energy conservation:
3118: \begin{align}
3119: L - \dot{\rho} \frac{\partial L}{\partial \dot{\rho}} = \text{const} \ ,
3120: \end{align}
3121: and one can see that the classical D1-brane is described by the trajectory
3122: \begin{align}
3123: \cosh \rho = \frac{\cosh r_0}{\cos \left[(\tilde{\theta} - \tilde{\theta}_0)/k\right]} \ .
3124: \end{align}
3125: The appearance of the $1/k$ in the argument of the cosine is important. If $k$ is an even integer, the asymptotic form of the D-brane trajectory is given by the coincident two-branes, while for an odd integer $k$, it is given by the parallel two-branes placed at the anti-podal points in $\mathbb{S}^1$.\footnote{For general $k$, asymptotic position of the two branes breaks the (discrete) periodic symmetry.}
3126: 
3127: The semiclassical boundary wavefunction is obtained by integrating the classical closed string wavefunction over the classical D-brane trajectory as
3128: \begin{align}
3129: \Psi_{2}(\tilde{\theta}_0,r_0)^{\mathrm{mini}} &= e^{iw\tilde{\theta}_0}\int_1^{\infty} \cosh\rho \dd(\cosh\rho) \int_{-\frac{k\pi}{2}}^{\frac{k\pi}{2}} \dd\tilde{\theta} \delta(\cos(\tilde{\theta}/k)\cosh\rho-\cosh r_0) \phi_{p,w}(\rho,\tilde{\theta}) \cr
3130: &= \int_{-\frac{k\pi}{2}}^{\frac{k\pi}{2}} \dd\tilde{\theta} \frac{\cosh r_0}{\cos^2(\tilde{\theta}/k)} \phi_{p,w}(\hat{\rho}(r_0,\tilde{\theta}),\tilde{\theta}) \ ,
3131: \end{align}
3132: where $\hat{\rho}(r_0,\tilde{\theta})$ is the solution of $\cos (\tilde{\theta}/k) \cosh \rho = \cosh r_0 $. The integration is feasible due to the formula
3133: \begin{align}
3134: \int_{-\frac{k\pi}{2}}^{\frac{k\pi}{2}} \dd\tilde{\theta} \frac{\cosh r_0}{\cos^2(\tilde{\theta}/k)} e^{iw\tilde{\theta}} (\cosh\rho)^{-1-ip} F\left(\frac{1}{2}-\frac{kw}{2}+\frac{ip}{2},\frac{1}{2}+\frac{kw}{2}+ \frac{ip}{2};1+ip;\frac{\cos^2(\tilde{\theta}/k)}{\cosh^2r_0}\right)\ \cr
3135: = \frac{2\pi \Gamma(ip)}{\Gamma(\frac{1}{2}+\frac{ip}{2}+\frac{kw}{2})\Gamma(\frac{1}{2}+\frac{ip}{2}-\frac{kw}{2})} e^{-ip r_0} \ , \label{evaluation overlap phi2}
3136: \end{align}
3137: whose derivation is relegated to the appendix B (see also \cite{Nakayama:2005pk}).
3138: 
3139: In this way, we derive the minisuperspace limit of the boundary wavefunction that describes the D1-brane in the trumpet geometry:
3140: \begin{align}
3141: \Psi_{2}(\tilde{\theta}_0,r_0)^{\mathrm{mini}} = N(b)\frac{\Gamma(2j+1)}{\Gamma(1+j+\frac{kw}{2})\Gamma(1+j-\frac{kw}{2})} e^{iw \tilde{\theta}_0} \cos(r_0(2j+1)) \ , \label{cltrum}
3142: \end{align}
3143: where $j = -\frac{1}{2} + \frac{ip}{2}$. This should be identified with the boundary wavefunction for the partially wrapped D2-brane via the T-duality. Note that the for zero winding sector $w=0$, the wavefunction agrees with the direct evaluation \eqref{direce}.
3144: 
3145: In a similar manner, the classical boundary wavefunction for the totally wrapped D2-brane is given by
3146: \begin{align}
3147: \Psi_{3}(\sigma,\theta_0) = \Gamma(2j+1)e^{i\theta_0\omega} \left[\frac{\Gamma(-j+\frac{kw}{2})}{\Gamma(j+1+\frac{kw}{2})}e^{i\sigma(2j+1)} + \frac{\Gamma(-j-\frac{kw}{2})}{\Gamma(j+1-\frac{kw}{2})}e^{-i\sigma(2j+1)} \right] \ . \label{clth}
3148: \end{align}
3149: The parameters $r_0$ and $\sigma$ are supposedly related to the magnetic flux on the D2-brane:
3150: \begin{align}
3151: F = \frac{\beta^2 \tanh^2\rho}{\cosh^2\rho - \beta^2} d\theta d\rho \ ,
3152: \end{align}
3153: where $\beta = \sin \sigma $ for $\beta \le 1$ and $\beta = \cosh r_0$ for $\beta \ge 1$. We expect that these two classes of branes coincide in the limit $\beta = 1$ (i.e. $\sigma = \pm \frac{\pi}{2}$ and $r_0 = 0$).
3154: 
3155: From the geometry of the semiclassical D-brane, we expect that if one takes a suitable limit of the boundary states, the class 2 D-brane (partially wrapped D-brane) will coincide with the class 3 D-brane (totally wrapped D-brane).
3156: To compare these two branes, we can directly show that 
3157: \begin{align}
3158: \Psi_3(\pi/2,-k\pi/2) + \Psi_3(\pi/2,k\pi/2) = \Psi_{2}(0,0) \ .
3159: \end{align}
3160: This result also shows that the boundary wavefunction \eqref{clth} describes the half cut D2-brane.
3161: 
3162: %\begin{align}
3163: %\psi_3(\pi/2) + \psi_3(-\pi/2) = \sin(\pi j) \left(\frac{1}{\sin\pi(j-\frac{kw}{2})} + \frac{1}{\sin\pi(j+\frac{kw}{2})}\right)\psi_2(0,0) \ .
3164: %\end{align}
3165: %Now we will see that when $k$ is an integer, it is possible to interpret (some combination of) the $r_0 \to 0$ limit of the class 2 branes as the class 3 brane (and its image under $w \to -w$ symmetry). 
3166: %\begin{itemize}
3167: %	\item When $k = 4n$ (with $n\in \mathrm{Z})$, there is nothing to say: $\psi_3(\pi/2) + \psi_3(-\pi/2) = \psi_2(0,0)$.
3168: %	\item When $k = 4n+2$, we have $\psi_3(\pi/2) + \psi_3(-\pi/2) = (-1)^w \psi_2(0,0) = \psi_2(0,\pi) $. These two cases are studied in \cite{Fotopoulos:2004ut}. 
3169: %	\item When $k=2n+1$, $\psi_3(\pi/2) + \psi_3(-\pi/2) = (-1)^{w/2}(1+(-1)^w)\psi_2(0,0) =  \psi_2(0,\pi/2) + \psi_2(0,3\pi/2)$.\footnote{Unfortunately this possibility seems inconsistent with the quantization condition for $\sigma$: $\sigma-\sigma' = \frac{2\pi m}{k}$ with an integer $m$.} 
3170: %\end{itemize}
3171: %We also note that in the zero-winding sector, the class 2 and the class 3 agree with each other consistent with the results discussed before.
3172: 
3173: %Therefore, we can conclude that up to a suitable choice of Wilson-line parameter $\theta_0$, we can connect class 2 brane and class 3 brane in this limit when $k$ is an integer. However, it is not clear whether it is possible to extend this relation for general (non-integer) $k$.\footnote{This is not unrelated to the fact that the class 3 brane satisfies the Cardy condition only when $k$ is an integer \cite{Israel:2005fn}.} 
3174: 
3175: %We have several comments:
3176: %\begin{itemize}
3177: %	\item The T-duality between the partially wrapped D2-brane in the cigar and the D1-brane in the trumpet is a little bit subtle when $k$ is not an integer. In this case, the asymptotic D1 brane breaks the (discrete) periodicity of the trumpet and the naive T-duality does not seem to apply.
3178: %	\item Nevertheless the class 2 brane is always consistent with the Cardy condition essentially because it was constructed via the modular bootstrap method. It should be contrasted with the class 3 boundary states which satisfies the Cardy condition only if one takes an integer value of $k$.\footnote{This will cause the failure of the coincidence between class 2 and class 3 brane in the $r_0 \to 0$ limit.} 
3179: %	\item If one restricts himself to the odd integer $k$, the class 2 brane seems to describe {\it two} partially wrapped D2-brane whose gauge symmetry $U(2)$ is spontaneously broken down to $U(1) \times U(1)$ due to the Wilson line. This interpretation is consistent with the $r_0 \to 0$ limit discussed in the last section.
3180: %	\item The class 2 boundary state has a symmetry under $w \to -w$ (for $\theta_0 = 0$) while the class 3 boundary state does not. As a consequence, in order to match these two boundary states, we have to symmetrize the class 3 boundary states by adding the contribution $\sigma \to -\sigma$. From the classical viewpoint, the signature of $\sigma$ is a matter of convention. However, the quantization condition: $\sigma-\sigma' = \frac{2\pi m}{k}$ will hinder this symmetrization even in the  $r_0 = 0$ limit. Thus, the smooth $r_0 \to 0$ limit in general cases is not so manifest.
3181: %	\item From the T-dual picture, it is very mysterious that the class 3 boundary state does not have a symmetry under $w \to - w$. Since in this regime, the D1-brane trajectory in the trumpet breaks up in two due to the existence of the singularity, it is likely that the class 3 brane only describes one part of the D1-brane. Still a mysterious thing is there should be a quantization condition $\sigma-\sigma' = \frac{2\pi m}{k}$. What is the origin?
3182: %	\item From the minisuperspace viewpoint, it is extremely puzzling that the class 2 (partially wrapped D2 in the cigar or connected D1 in the trumpet) does not reduce to the class 3 (totally wrapped D2 in the cigar or disconnected D1 in the trumpet) in a suitable limit. Since we have seen that the minisuperspace approximation is consistent with the class 2, we tentatively suspect that there should be something wrong with the minisuperspace analysis for the class 3. Due to the technical difficulty I have not been unable to do the minisuperspace integration for class 3 (except for zero-winding sector). It seems very interesting to complete this analysis.
3183: %\end{itemize}
3184: 
3185: %My tentative conclusions are
3186: %\begin{itemize}
3187: %	\item At least, from the semiclassical viewpoint, the class 2 D-brane describes correctly the D1-brane in the trumpet geometry. It is, however, not clear whether this is related to the partially wrapped D2-brane in the cigar geometry after T-duality, except for the case when $k$ is an integer.
3188: %	\item For the zero winding sector, minisuperspace analysis for the partially wrapped D2-brane is consistent with the class 2 D-brane.
3189: %	\item If the class 2 D-brane describes the partially wrapped D2-brane (and if the class 3 D-brane describes the totally wrapped D2-brane), there should  exist a suitable limit where these two branes will coincide. However this is not manifest because first of all, the class 2 D-brane seems to describe two (or more) D-branes with Wilson line, and on the other hand, it is difficult to combine the class 3 D-branes in an appropriate manner due to the quantization condition.
3190: %\end{itemize}
3191: 
3192: \subsubsection{embedding into NS5-branes}\label{sec:6-1-4}
3193: We have obtained the classical D-brane solutions in the (Euclidean) two-dimensional black hole background. As we have reviewed in section \ref{sec:2}, we can embed the two-dimensional black hole into the superstring theory as NS5-branes (or more generally little string theories on singular Calabi-Yau spaces). Here we would like to summarize some of the D-brane solutions in the NS5-brane background to see how one can construct them from those in the two-dimensional black hole system \cite{Elitzur:2000pq,Lerche:2000uy,Gava:2001gv,Eguchi:2003ik,Eguchi:2004yi,Eguchi:2004ik,Israel:2005fn}.
3194: 
3195: Let us first concentrate on the D1-brane solution in the ring-likely separated NS5-brane solution \eqref{rmet} corresponding to 
3196: \begin{align}
3197: \frac{\Big[ {SL(2;\br)_{k} \over U(1)} \times \frac{SU(2)_{k}}{U(1)} \Big]_\perp}{\bz_k} \ .
3198: \end{align}
3199: Naturally, we can combine various D-branes in the $SL(2;\br)/U(1)$ coset and $SU(2)/U(1)$ coset to construct D-branes in this background. We further focus on the two-plane $x^8=x^9 = 0$, setting $\theta =0$.
3200: 
3201: The first combination is the D0-brane in the cigar and the D1-brane in the bell. The result is the D1-brane stretching between NS5-branes as in figure \ref{fig:nsbranes}. In the context of the LST, we interpret them as W-bosons. 
3202: The second combination is the (uncut) D1-brane in the trumpet and D0-brane in the bell. The resulting geometry is the straight line on the $x^8=x^9=0$ plane as in figure \ref{fig:nsbranes}.
3203: The third combination is the cut D1-brane in the trumpet and D0-brane in the bell. It corresponds to the semi-infinite D-brane attached to the NS5-branes as in figure \ref{fig:nsbranes}.
3204: 
3205: The geometries of the D3-brane are much more complicated. We would like to refer to \cite{Ribault:2003sg,Israel:2005fn} for detailed study of the D3-brane geometries in the NS5-brane background.
3206: 
3207: Recently, a static D-brane configuration in the black hole background has attracted much attention for a possible application to the phase transition of the fundamental matters in QCD \cite{Mateos:2006nu}. In our two-dimensional black hole setup, it amounts to the study of the D-brane in the black NS5-brane background. In the Rindler limit studied in \cite{Mateos:2006nu}, the difference between the black NS5-brane and the black D-brane does not exist. It would be interesting to study the exact boundary states for the D-branes in the black NS5-brane background to probe the $\alpha'$ corrections to the phase transition discussed there.
3208: 
3209: 
3210: 
3211: \begin{figure}[htbp]
3212:    \begin{center}
3213:     \includegraphics[width=0.5\linewidth,keepaspectratio,clip]{nsbranes.eps}
3214:     \end{center}
3215:     \caption{The left figure shows W-bosons in LST. The central figure shows an uncut D1-brane. The right figure shows cut D1-branes attached to the NS5-branes.}
3216:     \label{fig:nsbranes}
3217: \end{figure}
3218: 
3219: \subsection{Exact boundary states}\label{sec:6-2}
3220: \subsubsection{Ishibashi states}\label{sec:6-2-1}
3221: To construct the exact Cardy boundary states for the D-branes in the two-dimensional black hole background, we begin with the Ishibashi states. For definiteness, we first concentrate on the bosonic axial coset, which is given by the Euclidean cigar.
3222: 
3223: The coset Ishibashi states naturally descend from those for the parent current algebra. The Ishibashi state satisfies the boundary condition
3224: \begin{align}
3225: (L_{n} - \bar{L}_{-n}) |A \rangle \rangle &= 0  \cr 
3226: (J_{n} -\bar{J}_{-n}) |A \rangle \rangle &= 0  \ ,
3227: \end{align}
3228:  for A-brane and
3229: \begin{align}
3230: (L_{n} - \bar{L}_{-n}) |B \rangle \rangle &= 0  \cr 
3231: (J_{n} + \bar{J}_{-n}) |B \rangle \rangle &= 0  \ ,
3232: \end{align}
3233:  for B-brane. In terms of the primary states of the coset, A-boundary condition means $m = \bar{m} = \frac{k\omega}{2}$, and B-boundary condition means $m = -\bar{m} = \frac{n}{2}$. Physically, the A-branes couple to the winding states while B-brane couple to the momentum states in the coset.
3234: 
3235: The Ishibashi state is naturally endowed with the classification via the character of the coset model. For continuous series, we have the following normalization 
3236: \begin{align}
3237: _B\langle \langle p',n'|e^{-\pi t(L_0 + \bar{L}_0)}| p,n \rangle \rangle_B = \left[\delta(p-p') + R(p,n)\delta(p+p')\right]\delta_{n,n'}\frac{q^{-\frac{p^2}{4(\kappa-2)}+\frac{n^2}{4k}}}{\eta(\tau)^2} \cr
3238: _A\langle \langle p',\omega'|e^{-\pi t(L_0 + \bar{L}_0)}| p,\omega \rangle \rangle_A = \left[\delta(p-p') + R(p,\omega)\delta(p+p')\right]\delta_{\omega,\omega'}\frac{q^{-\frac{p^2}{4(\kappa-2)}+\frac{\omega^2}{4}}}{\eta(\tau)^2} \ ,
3239: \end{align}
3240: where the subscript denote the boundary condition (either A-type or B-type), and $R(p,n)$ (or $R(p,\omega))$ denote the reflection amplitude. The Ishibashi state is parametrized by the radial momentum $p$ and the angular momentum $n$ (or the winding number $\omega$).
3241: 
3242: %The Ishibashi state corresponding to the discrete states $j=m = \frac{k\omega}{2}$ (or $j=-m$) have the following normalization
3243: %\begin{align}
3244: %_A\langle \langle p',\omega'|e^{-\pi t(L_0 + \bar{L}_0)}| p,\omega \rangle \rangle_A = \delta(p-p')\delta_{\omega,\omega'}\frac{q^{-\frac{p^2}{2}+\frac{\omega^2}{2k}}}{\eta(\tau)^2} \ . 
3245: %\end{align}
3246: %The discrete series only appear in the A-boundary condition because they need a winding quantum number. We also note that they should satisfy the (improved) unitarity constraint ($ -\frac{\kappa-1}{2}<j<-\frac{1}{2}$). 
3247: 
3248: For the supersymmetric coset, we impose the following boundary condition for the Ishibashi states:
3249: \begin{align}
3250: (L_n - \bar{L}_{-n}) | A \rangle \rangle = 0 \cr
3251: (G^{\pm}_r - i\bar{G}^{\mp}_{-r}) | A \rangle \rangle = 0 \cr
3252: (J_n - \bar{J}_{-n}) |A \rangle \rangle = 0 \ ,
3253: \end{align}
3254: for A-type boundary conditions, and
3255: \begin{align}
3256: (L_n - \bar{L}_{-n}) | B \rangle \rangle = 0 \cr
3257: (G^{\pm}_r - i\bar{G}^{\pm}_{-r}) | B \rangle \rangle = 0 \cr
3258: (J_n + \bar{J}_{-n}) |B \rangle \rangle = 0 \ ,
3259: \end{align}
3260: for B-type boundary conditions. Both types of the boundary conditions are compatible with the diagonal $\mathcal{N}=1$ superconformal symmetry
3261: \begin{align}
3262: (G_r - i \bar{G}_{-r})|A \ \mathrm{or} \ B\rangle\rangle \ ,
3263: \end{align}
3264: where $G_r = G^+_r + G^-_r$ that should be gauged in the fermionic string theory. Physically, A-type boundary condition corresponds to Dirichlet boundary condition along the cigar angular direction, and B-type boundary condition corresponds to Neumann boundary condition. 
3265: 
3266: The Ishibashi states for the supersymmetric coset for continuous series is parametrized by three quantum number $(p,m,s)$. Our normalization is
3267: \begin{align}
3268: _A\langle\langle p',\omega',s'| e^{-\pi\tau_c(L_0+\bar{L}_0)} e^{i\pi y(J_0+\bar{J_0})}|p,\omega,s\rangle \rangle_A \cr
3269: = \delta_{\omega',\omega}(\delta(p-p')+\delta(p+p')R(j,\frac{k\omega}{2},\frac{k\omega}{2})) \mathrm{ch}_{j,\frac{k\omega}{2},s}(i\tau_c,y) \ ,
3270: \end{align}
3271: for the A-brane, and 
3272: \begin{align}
3273: _B\langle\langle p',n',s'| e^{-\pi\tau_c(L_0+\bar{L}_0)} e^{i\pi y(J_0+\bar{J_0})}|p,n,s\rangle \rangle_B \cr
3274: = \delta_{n',n}(\delta(p-p')+\delta(p+p')R(j,\frac{n}{2},-\frac{n}{2})) \mathrm{ch}_{j,\frac{n}{2},s}(i\tau_c,y) \ ,
3275: \end{align}
3276: for the B-brane. Here $s$ denotes the spectral flow parameter. Note that the boundary condition demands $m=\bar{m}=\frac{k\omega}{2}$ for the A-brane and $m=-\bar{m}= \frac{n}{2}$ for the B-brane. The $\mathcal{N}=2$ character $\mathrm{ch}_{j,m,s}$ is defined as
3277: \begin{align}
3278: \mathrm{ch}_{j,m,s}(\tau,y) = q^{\frac{p^2}{4k}+\frac{(m+s)^2}{k}+\frac{s^2}{2}}z^{\frac{2m}{k}+s} \frac{\theta_3(\tau,y)}{\eta(\tau)^3} \ ,
3279: \end{align}
3280: for NS sector ($z=e^{2\pi i y}$).
3281: 
3282: \subsubsection{exact boundary wavefunction}\label{sec:6-2-2}
3283: Let us first summarize the exact boundary wavefunction for the D-branes whose classical properties we discussed in section \ref{sec:6-1}. We relegate a (partial) derivation of the exact boundary wavefunctions based on the modular bootstrap to section \ref{sec:6-2-3}.
3284: 
3285: For the bosonic two-dimensional black hole, we expand the Cardy boundary states
3286: as
3287: \begin{align}
3288: \langle B | = \int \dd p \sum_m \Psi(p,m) \langle\langle p,m | + (\text{discrete}) \ .
3289: \end{align}
3290:  Compared with the minisuperspace approximation \eqref{miniwv}, we have allowed the winding states (for B-brane) and a possible discrete state contribution. In the following, we focus on the continuous part. The discrete part can be read from the analytic continuation of the boundary wavefunction $\Psi(p,m)$ with respect to the parameters of the continuous series restricted to the value corresponding to the discrete series (i.e. $\Psi(j=m,m)$).
3291: 
3292: The exact boundary wavefunction for the D0-brane (class 1 A-type brane) is given by
3293: \begin{align}
3294: \Psi_{\mathrm{D}0}(j,\omega) = \nu_b^{2j+1} \frac{\Gamma(-j+\frac{k\omega}{2})\Gamma(-j-\frac{k\omega}{2})}{\Gamma(-2j-1)\Gamma(1-b^2(2j+1))} \ , \label{exco}
3295: \end{align}
3296: where $b = (k-2)^{-1/2}$, and $\nu_b = \frac{\Gamma(1-b^2)}{\Gamma(1+b^2)}$. It is easy to see that the exact boundary wavefunction \eqref{exco} reduces to the mini-superspace result \eqref{minico} in the large $k$ limit by setting $\omega = 0$ up to a $p$ independent overall normalization factor. The exact boundary state for the D0-brane couples to winding states. It also couples to the discrete series localized near the tip of the cigar.
3297: 
3298: The exact boundary wavefunction for the D1-brane (class 2' B-type brane) is given by
3299: \begin{align}
3300: \Psi_{\mathrm{D}1}(j,n)^{r,\theta_0} =   \nu_b^{2j+1} e^{in\theta_0} \frac{\Gamma(2j+1)\Gamma(1+b^2(2j+1))}{\Gamma(1+j+\frac{n}{2})\Gamma(1+j-\frac{n}{2})} (e^{-r(2j+1)}+(-1)^n e^{r(2j+1)}) \ . \label{clss2'}
3301: \end{align}
3302: The D1-brane only couples to the momentum states. In particular, it does not couple to any discrete states. In the classical limit $k\to \infty$, the boundary wavefunction reproduces that of the minisuperspace approximation \eqref{hairpin
3303: D1 classical}.
3304: 
3305: The exact boundary wavefunction for the partially wrapped D2-brane (class 2 A-type brane) is given by
3306: \begin{align}
3307: \Psi_{\mathrm{D}2}(j,\omega)^{r_0,\tilde{\theta}_0} = \nu_b^{2j+1} \frac{\Gamma(2j+1) \Gamma(1+b^2(2j+1))}{\Gamma(1+j+\frac{kw}{2})\Gamma(1+j-\frac{kw}{2})} e^{iw \tilde{\theta}_0} \cos(r_0(2j+1)) \ . \label{exbt}
3308: \end{align}
3309: It does not couple to the discrete states localized at the tip of the cigar as is expected from the geometry. We can readily see that the classical limit ($k\to \infty$) of \eqref{exbt} reduces to the minisuperspace wavefunction \eqref{cltrum}. 
3310: 
3311: Finally, the exact boundary wavefunction of the totally wrapped D2-brane (class 3 A-type brane) is given by
3312: \begin{align}
3313: \Psi_{\mathrm{D}2'}(j,\omega)^{\sigma,{\theta}_0} &= \nu_b^{2j+1} \Gamma(1+b^2(2j+1))\Gamma(2j+1)e^{i\theta_0\omega}  \times \cr 
3314: &\times \left[\frac{\Gamma(-j+\frac{kw}{2})}{\Gamma(j+1+\frac{kw}{2})}e^{i\sigma(2j+1)} + \frac{\Gamma(-j-\frac{kw}{2})}{\Gamma(j+1-\frac{kw}{2})}e^{-i\sigma(2j+1)} \right] \ ,\label{exbthr}
3315: \end{align}
3316: with the relative quantization condition $\sigma-\sigma' = 2\pi\frac{m}{k-2}$ , $m \in \bz$.
3317: 
3318: The other possible exact boundary states for the two-dimensional black hole have been proposed in the literatures \cite{Ahn:2004qb,Hosomichi:2004ph,Ribault:2005pq}. However, they do not possess sensible open string spectra\footnote{Most of them contain tachyon in their spectra. Furthermore, they often have imaginary conformal weights when we study overlaps with class 2 (class 2') branes.} nor corresponding semiclassical limits, so we will not discuss them in detail here. Their properties are in many sense similar to the generalized ZZ branes proposed in \cite{Zamolodchikov:2001ah}. As such they could be important so as to understand the nonperturbative contributions to the partition function of the two-dimensional Euclidean black hole.
3319: 
3320: The boundary wavefunctions for the supersymmetric two-dimensional black hole are essentially the same as those for the bosonic one. In the NS sector, the only difference is to replace $b^2 = \frac{1}{k-2}$ with $\frac{1}{k}$. The boundary wavefunction for the other sector is obtained by the spectral flow.
3321: \subsubsection{Cardy condition and modular bootstrap}\label{sec:6-2-3}
3322: There are several different ways to derive the boundary wavefunctions for the D-branes in the $SL(2;\br)/U(1)$ coset model. One of the simplest ways to obtain them is to descend them from the branes in the parent $SL(2;\br)$  WZNW model (or $\mathbb{H}^3_+$ model). This method has a small drawback for our purposes because we should derive the boundary states for $SL(2;\br)$ WZNW model (or $\mathbb{H}_3^+$ model) first \cite{Ponsot:2001gt}. In this section, we take another root, which uses the so-called ``modular bootstrap" method to derive all the A-branes in the Euclidean two-dimensional black hole.\footnote{As we will see, we cannot derive the boundary wavefunctions for B-branes in this approach. We need a more technically involved strategy such as the conformal bootstrap to derive them.}
3323: 
3324: The modular bootstrap method is intimately related to the Cardy condition for boundary states \cite{Cardy:1989ir}. The Cardy condition is the physical constraint on the open string spectra between two different D-branes. Let us denote the boundary states for any pair of these two branes as $|a\rangle$ and $|b\rangle$. The Cardy condition says that the open string spectra between these two D-branes should have open string characters with positive multiplicities:
3325: \begin{align}
3326: Z_{a,b} = \mathrm{Tr}_{a,b} q^{L_o} = \sum_i n^i_{a,b}\chi_i(q) \ , \label{cardyc}
3327: \end{align}
3328: where $\chi_i(q)$ is the open string character, and $n^i_{a,b}$ should be positive integers from the unitarity of the theory.\footnote{Implicitly here we are assuming that the open string between $|a\rangle$ and $|b\rangle$ are bosonic. Otherwise the negative multiplicity is allowed as fermions. For NS-NS overlap, we expect that the overlap should contain bosonic excitations.} Now we modular transform the open string character $\tau \to -1/\tau$ in order to obtain the closed string description:
3329: \begin{align}
3330: \langle a|e^{i\tau \pi H_c}| b\rangle = Z_{ab}(q) = \sum_{i,j} n^{i}_{a,b}S_{ij} \chi_j(\tilde{q}) \ ,
3331: \end{align}
3332: with $\tilde{q} \equiv e^{-2\pi i/\tau}$, where $S_{ij}$ is the modular $S$-matrix for characters $\chi_i$. 
3333: 
3334: At this point, it is not immediately obvious whether the multiplicities $n^{i}_{a,b}$ are all positive integers if one introduces an arbitrary set of boundary states $|a\rangle$. This integrality condition is the Cardy condition for boundary states. The Cardy condition guarantees a physical interpretation of the open string spectra from the open-closed duality.
3335: 
3336: Actually, there is a canonical solution of the Cardy condition based on the Verlinde formula \cite{Verlinde:1988sn}. We assume that the boundary state is a superposition of the Ishibashi states with normalization
3337: \begin{align}
3338: \langle\langle i|e^{i\tau \pi H_c}|j\rangle\rangle = \delta_{ij}\chi_i(q) \ .
3339: \end{align}
3340: We then assume the existence of the simplest boundary states (identity brane) $|\hat{0}\rangle$ whose self overlap gives an identity representation in the open string sector: $n^i_{\hat{0},\hat{0}} = \delta^i_0$. Since the fusion of the identity operator is itself, we have a relation
3341: \begin{align}
3342: |\langle \hat{0}|j\rangle|^2 = S_{0,j} \ . 
3343: \end{align}
3344: As a consequence, the state $|\hat{0}\rangle$ can be written as
3345: \begin{align}
3346: |\hat{0}\rangle = \sum_j\sqrt{S_{0,j}}|j\rangle\rangle \ , \label{sob}
3347: \end{align}
3348: up to an overall phase factor (possibly dependent on $j$). 
3349: 
3350: Now we {\it define} Cardy boundary states as
3351: \begin{align}
3352: |a\rangle = \sum_j \frac{S_{aj}}{\sqrt{S_{0j}}}|j \rangle\rangle \label{sobb}
3353: \end{align}
3354: which contains the open string spectrum $n^{i}_{\hat{0}a} = \delta_a^i $ in the overlap with the identity brane. These Cardy states satisfy the Cardy condition
3355:  \eqref{cardyc} 
3356: \begin{align}
3357: \langle a|j \rangle \langle j| b \rangle = \frac{S_{aj}S_{jb}}{S_{0j}} = \sum_i S_{ij} n^{i}_{ab} \ ,
3358: \end{align}
3359: where $n^{i}_{ab}$ is given by the Fusion coefficient $\mathcal{N}^{i}_{ab}$ that is a positive integral matrix. The last equality is due to a remarkable identity under the name of the Verlinde formula. The Verlinde formula can be shown for unitary compact CFTs by studying the monodromy constraint for the torus amplitudes.
3360: 
3361: Let us move on to the boundary wavefunction for A-branes in the two-dimensional black hole. Our first assumption is the existence of the identity brane, which will be identified as the D0-brane at the tip of the cigar. We assume that the self-overlap of this identity brane yields the identity representation summed over the spectral flow in the open string spectrum. Summation over the spectral flow is needed in order to guarantee the quantization of the $U(1)_R$ charge in the closed string spectrum.\footnote{Otherwise, we obtain the D-brane in the decompactified theory, which has been studied in \cite{Ahn:2003tt,Ahn:2004qb}, in the context of the $\mathcal{N}=2$ Liouville theory. We also note that our summation over the spectral flow is different from \cite{Eguchi:2003ik}, where the summation is taken over $n\in k \mathbb{Z}$ for integral $k$. For A-brane, the latter summation leads to the integral $U(1)_R$ charge (i.e. fractional $\omega$ quantum number).}
3362: 
3363: For definiteness, we study the NS-sector of the supersymmetric $SL(2;\br)/U(1)$ coset. Our assumption mentioned above is 
3364: \begin{align}
3365: Z_{00} = \langle 0|e^{-\pi\tau_c(L_0+\bar{L}_0 -\frac{c}{12})+iy(J_0+\bar{J}_0)}|0\rangle = \sum_n \frac{\mathrm{ch}_{0,n,n}(\tau_o)(1-q_o)}{(1+yq_o^{\frac{1}{2}+n})(1+y^{-1}q_o^{\frac{1}{2}-n})} \ . \label{ajka}
3366: \end{align}
3367: The modular S-transformation of the extended character of the identity representation (the identity character summed over the spectral flow) is given by \cite{Eguchi:2003ik}: 
3368: \begin{align}
3369: Z_{00} = \int_{-\frac{1}{2}+i\br_+} \dd j \sum_{m\in \frac{k}{2}Z} \frac{i\sin(\pi(2j+1))\sin\frac{\pi}{k}(2j+1)}{2\sin(j+m)\pi\sin(j-m)\pi} \mathrm{ch}_{j,m,0}(\tau_c) + \text{(discrete)} \ . \label{idmod}
3370: \end{align}
3371: The discrete terms are a little bit trickier to obtain, and we refer the complete form to original papers \cite{Eguchi:2003ik}.
3372: 
3373: The boundary wavefunction is essentially obtained by taking the square root of the modular S-matrix in analogy with \eqref{sob}. Expanding the identity boundary state as
3374: \begin{align}
3375: \langle 0|= \int_{-\frac{1}{2}+i\br_+} \dd j \sum_m \Psi(j,m)_0 \langle\langle j,m|\ ,
3376: \end{align}
3377: we obtain
3378: \begin{align}
3379: \Psi(j,m)_0 = \nu_b^{2j+1} \frac{\Gamma(-j+\frac{k\omega}{2})\Gamma(-j-\frac{k\omega}{2})}{\Gamma(-2j-1)\Gamma(1-b^2(2j+1))} \ .
3380: \end{align}
3381: We should note that the condition does not determine the $(j,m)$ dependent phase factor of the boundary wavefunction. The ambiguity of the phase, however, is completely fixed by the reflection relation
3382: \begin{align}
3383: \Psi(-p,m) = R(p,m) \Psi(p,m) \ ,
3384: \end{align}
3385: together with the Hermiticity condition $\Psi(p,m)^\dagger = \Psi(-p,-m)$. 
3386: 
3387: The next assumption is the overlap between the identity brane $|0\rangle$ and the general brane $|a \rangle$ labeled by the character of the $SL(2;\br)/U(1)$ coset model in analogy with \eqref{sobb}:
3388: \begin{align}
3389: \langle 0|e^{-\pi\tau_c(L_0+\bar{L}_0 -\frac{c}{12})+iy(J_0+\bar{J}_0)}|a\rangle = \chi_a(i\tau_o,y) \ . \label{anz}
3390: \end{align}
3391: We assume that the open string character appearing in the right hand side is given by the continuous series summed over the spectral flow (class 2 brane) and the discrete series summed over the spectral flow (class 3 brane). 
3392: 
3393: The S-modular transformation of the extended character for the continuous series (parametrized by $J$ and $M$) is particularly easy:
3394: \begin{align}
3395: \sum_{n\in \bz} \mathrm{ch}_{J,M+n,n}(\tau_o,y) = -i \sum_{m\in \frac{k}{2}Z} \int_{-\frac{1}{2}+i\br_+} \dd j \mathrm{ch}_{j,m,0}(\tau_c,y)e^{-\frac{4\pi iMm}{k}}\cos[\frac{\pi}{k}(2j+1)(2J+1)] \ .
3396: \end{align}
3397: From the modular bootstrap ansatz \eqref{anz}, we obtain the boundary wavefunction corresponding to the continuous series as
3398: \begin{align}
3399: \Psi(p,m)_{J,M} &= \Psi(-p,-m)_{0}^{-1} e^{-\frac{4\pi iMm}{k}} \cos[\frac{\pi}{k}(2j+1)(2J+1)] \cr
3400: &= \nu_b^{2j+1}\frac{\Gamma(2j+1)\Gamma(1+b^2(2j+1))}{\Gamma(1+j+\frac{kw}{2})\Gamma(1+j-\frac{kw}{2})} e^{-\frac{4\pi iMm}{k}}\cos[\frac{\pi}{k}(2j+1)(2J+1)] \ .  
3401: \end{align}
3402: With the parameter identification $ r_0 = \frac{\pi}{k}(2J+1)$, $\theta_0 = -2\pi M$, we have obtained the boundary wavefunction for the partially wrapped D2-brane.
3403: 
3404: The S-modular transformation of the extended character for the discrete series are more involved:
3405: \begin{align}
3406: &\sum_{n\in \bz} \frac{y^{-\frac{2}{k}(M+n)}\mathrm{ch}_{J,M+n,n}(\tau_o,y)}{1+y^{-1}q_o^{\frac{1}{2}+J-M-n}}  \cr
3407: &= -i \sum_{m\in \frac{k}{2}Z} \int_{-\frac{1}{2}+\br_+} \dd j \mathrm{ch}_{j,m}(\tau_c,y) e^{2\pi i M \omega} \left[\frac{e^{i(2J+1) (2j+1)}}{\sin\pi(j-\frac{kw}{2})} - \frac{e^{-(2J+1)(2j+1)}}{\sin(\pi(j+\frac{kw}{2}))}\right] \cr &+ \text{(discrete)} \ .
3408: \end{align}
3409: From the modular bootstrap ansatz \eqref{anz}, we obtain the boundary wavefunction corresponding to the discrete series as
3410: \begin{align}
3411: &\Psi(p,m)_{J,M} \cr
3412: &= \nu_b^{2j+1}\Gamma(1+b^2(2j+1))\Gamma(2j+1)e^{2\pi i M \omega}  \times \cr 
3413: &\times \left[\frac{\Gamma(-j+\frac{kw}{2})}{\Gamma(j+1+\frac{kw}{2})}e^{i\pi(m-j)+i(2J+1)(2j+1)} - \frac{\Gamma(-j-\frac{kw}{2})}{\Gamma(j+1-\frac{kw}{2})}e^{i\pi(m+j)-i(2J+1)(2j+1)} \right]  \ ,
3414: \end{align}
3415: where the parameter identification with the class 3 brane is $\theta_0 = 2\pi M + \frac{k\pi}{2} $ , and $\sigma = -\frac{\pi}{2} + \frac{\pi}{k}(2J+1)$.
3416: 
3417: So far, we have obtained as many branes as the extended character of the $\mathcal{N}=2$ supersymmetry (or $SL(2;\br)/U(1)$ coset). We, however, have to check whether the obtained D-branes satisfy the Cardy condition among themselves. Namely, we have to compute the cylinder amplitudes and decompose them into the open string characters and verify the positive definiteness of the density of states for continuous series and the positive integral multiplications for the discrete series. This was automatically guaranteed for the compact unitary CFTs thanks to the Verlinde formula. In our noncompact case, it is not a trivial problem. Indeed, although, almost all overlaps are consistent with the Cardy condition, the self-overlaps between class 3 branes for irrational value of $k$, we encounter negative multiplicities of discrete series in their spectra \cite{Ribault:2003ss}.\footnote{For integral value of $k$, this subtlety is avoided \cite{Israel:2005fn}. For general fractional value of $k$, the situation is more involved and the results depend on the combination of the other sectors embedded in the full string theory and the appropriate GSO condition we impose. The case by case study of these cases can be found in \cite{Eguchi:2003ik}.}
3418: 
3419: 
3420: One might wonder what goes wrong with the modular bootstrap for the B-branes. The gist is that there is no identity brane for the B-boundary conditions. One can formally write down the modular bootstrap equations like \eqref{idmod}, but there does not exist any analytic solution compatible with the reflection amplitudes for B boundary conditions. Due to this lack of the identity B-brane, the whole construction of the modular bootstrap breaks down.
3421: The coset construction from the descent of branes in $\mathbb{H}_3^+$ model was given in  \cite{Ribault:2003ss}. The conformal bootstrap for the dual $\mathcal{N}=2$ Liouville theory can be found in \cite{Ahn:2003tt,Hosomichi:2004ph}
3422: 
3423: \newpage
3424: \sectiono{Rolling D-brane in Two-dimensional Black Hole}\label{sec:7}
3425: In this section, we study the D-branes in Lorentzian two-dimensional black hole. The organization of the section is as follows. In section \ref{sec:7-1}, we study the classical D-branes in the Lorentzian two-dimensional black hole. In section \ref{sec:7-2}, we construct the boundary states for the rolling D-brane from the Wick rotation of the class 2 brane in the Euclidean two-dimensional black hole system.\footnote{This part of the thesis is based on \cite{Nakayama:2005pk}.} In section \ref{sec:7-3}, we study some properties of our boundary wavefunction focusing on $1/k$ corrections.
3426: \subsection{Classical D-branes}\label{sec:7-1}
3427: \subsubsection{DBI analysis}\label{sec:7-1-1}
3428: The classical D-branes in Lorentzian two-dimensional black hole
3429:  is classified by the solution of the equation of motion coming from the DBI action
3430: \begin{align}
3431: S^L = \mu_{p+1} \int \dd^{p+1}\xi e^{-\Phi} \sqrt{-\det(G_{ab}+B_{ab}+F_{ab})} \ . 
3432: \end{align}
3433: The classical background is given by
3434: \begin{align}
3435: \dd s^2 =k\al'(  - \tanh^2\rho \, \dd t^2 +  \dd
3436: \rho^2) ,
3437: \qquad e^{2\Phi} = \frac{k}{\mu \cosh^2 \rho} \ ,
3438: \end{align}
3439: or when we are interested in the global structure of the solution, we use the Kruscal coordinate
3440: \begin{align}
3441: \dd s^2 = -2k\frac{\dd u\dd v}{1-uv} \ , \ \ e^{2\Phi} = \frac{k}{\mu(1-uv)} \ . \label{krscc}
3442: \end{align}
3443: by the coordinate transformation: $u = \sinh \rho e^{t}$, $v=-\sinh\rho e^{-t}$.
3444: 
3445: We begin with the D(-1) instanton. A physical meaning of such D-brane is a little bit unclear in the Lorentzian signature, but the ``effective action"
3446: \begin{align}
3447: S_{-1} \propto e^{-\phi} = \sqrt{1-uv}  
3448: \end{align}
3449: could be extremized at $u=v=0$ or $\rho = 0$. 
3450: 
3451: Next we study the D0-brane. In the local coordinate outside the horizon, we can write down the DBI action as
3452: \begin{align}
3453: S_0 = \mu_0 \int \dd t \cosh\rho(t)\sqrt{-\dot{\rho}(t)^2 + \tanh^2\rho(t)} \ , \label{dbidzero}
3454: \end{align}
3455: where we have fixed the reparametrization invariance by taking the temporal gauge $\xi_0 = t$. From the energy conservation, we obtain
3456: \begin{align}
3457: \text{const} = \cosh\rho\frac{\tanh^2\rho}{\sqrt{-\dot{\rho}^2 + \tanh^2\rho}} \ ,
3458: \end{align}
3459: which can be integrated to
3460: \begin{align}
3461: \sinh(\rho) \cosh(t-t_0) = \text{const} \ . \label{mot}
3462: \end{align}
3463: The D0-brane motion \eqref{mot} also follows from the Wick rotation $\theta \to it$ to the hairpin brane \eqref{trsce} in the Euclidean two-dimensional black hole. As we mentioned in section \ref{sec:5-3}, The action \eqref{dbidzero} can be rewritten in the same form as the rolling D-brane in the linear dilaton background by
3464: introducing `tachyon' variable $Y \equiv \log \sinh \rho$:
3465: \begin{align} L_{\rm D0} = - V(Y) \sqrt{1 - \dot{Y}^2} \qquad \mbox{where}
3466: \qquad
3467:  V(Y) = M_0 \, e^Y ~, \label{nonextremal D0} \end{align}
3468: which leads us to the ``tachyon - radion correspondence" discussed in section \ref{sec:5}.
3469: 
3470: To study the global structure, we use the Kruscal coordinate \eqref{krscc}. In this coordinate system, the DBI action takes the flat form
3471: \begin{align}
3472: S = \mu_0  \int \dd\xi \sqrt{\frac{\dd u}{\dd \xi}\frac{\dd v}{\dd\xi}} \ .
3473: \end{align}
3474: The equation of motion is solved by a straight line in the $(u,v)$ plane. It is interesting to note that the D0-brane does not feel the existence of the singularity at $uv=1$. The classical trajectory is analytically continued inside the singularity in a trivial way. This is because the curvature singularity is cancelled against the dilaton singularity, which appears in the DBI action in the opposite way. The coupling to the dilaton is a crucial difference between the D-brane and a usual particle (such as a point like F-string or folded string solution discussed in section \ref{sec:3-3-2}) in the two-dimensional black hole background. 
3475: 
3476: Let us finally consider the D1-brane. The DBI action in the Kruscal coordinate is
3477: \begin{align}
3478: S_1 = \mu_1 \int \dd u\dd v \sqrt{1-uv}\sqrt{\frac{1}{(1-uv)^2} - F_{uv}^2} \  
3479: \end{align}
3480: in the gauge $\xi_0 = u$, $\xi_1 = v$. The Gauss law constraint
3481: \begin{align}
3482: f = \frac{\sqrt{1-uv}F_{uv}}{\sqrt{\frac{1}{(1-uv)^2}-F^2_{uv}}} 
3483: \end{align}
3484: is solved by
3485: \begin{align}
3486: F_{uv}^2 = \frac{f^2}{(1-uv+f^2)(1-uv)^2} \ .
3487: \end{align}
3488: When $f^2>0$, the world-sheet of the D1-string covers the whole physical region of the two-dimensional black hole, and possibly it has a boundary inside the singularity at $uv=1+f^2$. When $f^2 = -\kappa^2 <0$, the D1-string has a boundary at $uv = 1-\kappa^2$, and describes a long folded string. However, the DBI action for the long folded string becomes imaginary, so the solution is overcritical and unphysical.\footnote{This is a general feature of the Lorentzian solution for the D-brane with the boundary coming from the blow-up of the field strength. In the Lorentzian signature, the terms inside the square root of the DBI action is bounded, or in other words the field strength has a critical value. Thus any D-brane that has a boundary due to the divergence of the field strength is overcritical and hence unphysical unlike the case in the Euclidean signature.}
3489: \subsubsection{group theoretical viewpoint}\label{sec:7-1-2}
3490: As we have done in section \ref{sec:6-1-2}, we can also study the classical D-brane in the Lorentzian two-dimensional black hole from the coset construction. We parametrize the parent $SL(2;\br)$ element $g$ as
3491: \begin{align}
3492: g = \begin{pmatrix} a \ u \cr -v \ b \end{pmatrix} \ , \ \ uv + ab = 1 \ .
3493: \end{align}
3494: Under the axial gauge transformation $\delta g = \epsilon(\sigma_3 g + g \sigma_3) $, the $(u,v)$ is invariant, and serves as a gauge invariant coordinate describing the two-dimensional black hole (i.e. we can identify them as $(u,v)$ in the Kruscal coordinate \eqref{krscc}).
3495:  
3496: The maximally symmetric D-brane in the parent $SL(2;\br)$ WZNW model is classified by the (twined) conjugacy class of the group. The conjugacy class is given by
3497: \begin{align}
3498: \mathrm{Tr}(g) = a + b 
3499: \end{align}
3500: and the twined conjugacy class is given by
3501: \begin{align}
3502: \mathrm{Tr}(\sigma_1 g) = u-v \ , \label{twcl}
3503: \end{align}
3504: up to a conjugation (i.e. Lorentz boost between $\sigma_1$ and $\sigma_2$). 
3505: To derive the D-branes in the coset model, we have to project the twined conjugacy class to the coset variables for A-branes. For B-branes, we first take a superposition of the gauge orbit of the conjugacy class so that we obtain a gauge invariant object as a D-brane. 
3506: 
3507: Let us begin with the A-branes. The twined conjugacy class \eqref{twcl} is invariant under the axial gauge transformation, so the D-brane (D0-brane) is classified by the equation
3508: \begin{align}
3509:  2\kappa = \mathrm{Tr}(\sigma_1 g ) = u- v \ ,
3510: \end{align}
3511: which gives a straight line in the Kruscal coordinate of the two-dimensional black hole as we observed in section \ref{sec:7-1-1} from the DBI analysis. More general branes are obtained by the Lorentz boost: 
3512: \begin{align}
3513: ue ^{t_0} - ve^{-t_0} = 2\kappa \ .
3514: \end{align}
3515: The existence of such boosted D-branes are consistent with the fact that the existence of the Nambu-Goldstone modes associated with the symmetry breaking due to the A-brane as we have reviewed in section \ref{sec:6-1-2}.
3516: 
3517: Let us move on to the B-branes. In this case, the conjugacy class is not invariant under the axial gauge transformation. In order to obtain an invariant object that can be projected down to the coset, we need to sum over the gauge orbit. With fixing the trace as $a+b = 2\kappa$, the determinant constraint reads
3518: \begin{align}
3519: uv = 1-\kappa^2 + (a-\kappa)^2 \ge 1-\kappa^2 \ .
3520: \end{align}
3521: It is not difficult to see that the gauge orbit of the conjugacy class precisely agrees with the domain bounded by the last inequality. The string configuration reproduces the folded D1-string obtained from the DBI analysis in section \ref{sec:7-1-1}. We note, however, that the solution is overcritical and unphysical as we have seen in section \ref{sec:7-1-1}.\footnote{It is not uncommon that the group theoretical classification of the D-branes in the Lorentzian coset gives unphysical D-branes (see e.g. \cite{Hikida:2005vd}). Our identification of the parameter is different from the one given in \cite{Yogendran:2004dm}, which solves a small puzzle raised there. The extra $i$ comes from a (hypothetical) time-like T-duality \cite{Hull:1998vg} which we need to perform to obtain the parameter identification according to the discussion given in \cite{Walton:2002db}. } We also see that the D-string that covers the whole physical region of the two-dimensional black hole cannot be obtained from a simple descent from the D-branes in the $SL(2;\br)$ WZNW model without an analytic continuation.
3522: 
3523: \subsection{Boundary states from Wick rotation}\label{sec:7-2}
3524: 
3525: \subsubsection{analytic continuation of boundary states}\label{sec:7-2-1}
3526: 
3527: In this section, we shall construct the exact boundary state
3528: describing the D0-brane moving in the Lorentzian two-dimensional
3529: black hole background. Recall that the Lorentzian two-dimensional
3530: black hole (`Lorentzian cigar') background is obtainable by the Wick
3531: rotation $\theta = it$ of the Euclidean one \eqref{Euclidean cigar}
3532: \begin{equation}
3533: \dd s^2 = 2k(\dd \rho^2 - \tanh^2\! \rho  \, \dd t^2) \qquad
3534: \mbox{and} \qquad e^{\Phi} = \frac{e^{\Phi_0}}{\cosh\rho} ~.
3535: \label{Lorentzian cigar}
3536: \end{equation}
3537: Wick-rotating the geodesic of the Euclidean D1-brane, we found the
3538: geodesic of the Lorentzian D0-brane in \eqref{mot} as
3539: %\footnote
3540: %{Another familiar parametrization of the two-dimensional
3541: %lack hole is the analogue of the Kruscal coordinates
3542: %$$
3543: %u= \sinh \rho e^t ~, ~~~ v=-\sinh \rho e^{-t} ~, ~~~ \dd s^2 = -{2k}
3544: %\frac{\dd u \dd v}{1-uv}~,
3545: %$$
3546: %  and the geodesic \eqref{trajectory D0} is just a straight line
3547: %  in these coordinates. This is also pointed out in \cite{Yogendran}.
3548: %}
3549: \begin{align}
3550:  \cosh(t-t_0) \sinh \rho = \sinh \rho_0~,
3551: \label{trajectory D0}
3552: \end{align}
3553: where $t_0$, $\rho_0$ are free parameters. Notice that the D0-brane
3554: reaches the horizon $\rho = 0$ at $t \rightarrow \pm \infty$
3555: irrespective of the values of $\rho_0$ and $t_0$. Thus, formally, the
3556: Lorentzian D0-brane boundary state is obtainable by Wick rotation of
3557: the Euclidean D1-brane boundary state \eqref{clss2'} if we are interested in the physics outside the event horizon.\footnote{Some
3558: classical analysis of D-brane dynamics was attempted in
3559: \cite{Yogendran:2004dm} within the Dirac-Born-Infeld approach.}
3560: 
3561: Reconstructing boundary states of the Lorentzian D-brane from those
3562: of the Euclidean D-brane is generically not unique. Rather, the
3563: following potential subtleties need to be faced:
3564: \begin{itemize}
3565:  \item The Euclidean momentum $n$ along the asymptotic circle
3566: of cigar is quantized, while the corresponding quantum number in the
3567: Lorentzian theory ({\em i.e.} the energy) takes a continuous value.
3568: %%%
3569:  \item The Wick rotations of primary states are not necessarily
3570: unique. Often, appropriate boundary conditions should be specified.
3571: \end{itemize}
3572: As for the first point, which has to do with Matsubara formulation,
3573: we can formally avoid the difficulty of quantized momentum by the
3574: following heuristic consideration. Suppose the boundary wave
3575: function $\hat{f}(n,\al)$ ($n\in \bz$ is the quantized Euclidean
3576: energy, and $\al$ denotes the remaining quantum numbers not touched
3577: here) is given by the Fourier transform of a periodic function
3578: $f(x+2\pi, \al)=f(x, \al)$. We then obtain
3579: \begin{align}
3580: \hspace{-1cm} \bra{B}=
3581: \sum_{\al}\sum_{n\in\bsz}\,\widetilde{f}(n,\al)\dbra{n,\al} &=
3582: \sum_{\al}\sum_{n\in\bsz}\, \frac{1}{2\pi}\int_{-\pi}^{\pi} \dd x\,
3583: f(x,\al) e^{in x} \, \dbra{n,\al} \nn &=  \sum_{\al}
3584: \int_{-\infty}^{\infty} \frac{\dd q}{2\pi}\, \int_{-\infty}^{\infty}
3585: \dd x\, f(x,\al) e^{iqx} \, \dbra{q,\al}~, \label{formal extension}
3586: \end{align}
3587: where we used the identity $ \sum_{n\in \bsz}\, \delta(q-n) =
3588: \sum_{m\in \bsz}\, e^{2\pi i m q} $ in obtaining the last
3589: expression. Assuming that $f(x, \al)$ is analytic along the entire
3590: real $x$ axis, the Wick rotation can be performed. Often, $f(x,
3591: \al)$ is non-analytic over the real $x$ axis, and the integral in
3592: the last expression is ill-defined. This turns out to be the case
3593: for the boundary wave function of the Euclidean D1-brane
3594: \eqref{clss2'}: in the coordinate space, the wave function has
3595: branch cuts and singularities along the real $x$-axis. In such
3596: cases, the best we can do is to adopt the slightly deformed
3597: integration contour $\cC$ in $x$-space\footnote
3598:    {To be more precise, we should allow to use
3599:   some decomposition
3600: $$
3601:  f(x,\al) = f_1(x,\al)+ f_2(x,\al)+\cdots~,
3602: $$
3603: and to take the different contours for each piece $f_i(x,\al)$. } to
3604: render the Fourier integral well-defined:
3605: \begin{align}
3606: && \bra{B'} \Big\vert_{\rm Euclidean} :=
3607: \sum_{\al}\int_{-\infty}^{\infty} \frac{\dd q}{2\pi}\, \int_{\cC}
3608: \dd x\, f(x,\al) \, e^{iqx} \, \dbra{q,\al} ~. \label{formal
3609: extension 2}
3610: \end{align}
3611: Likewise, disk one-point function of vertex operator $\Phi^{\rm
3612: Euclidean}_{q,\al}$ (associated with the Ishibashi state
3613: $\dbra{q,\al}$) is evaluated as the deformed contour integral:
3614: \begin{align}
3615: & \Big< \Phi^{\rm Euclidean}_{q,\al} \Big>_{\msc{disk}}
3616:    = {}_{\rm E}\!\langle B' \vert q, \al \rangle\rangle = \int_{\cC} \dd x\, f(x,\al) e^{iqx}~.
3617: \label{formal disk amp 1}
3618: \end{align}
3619: Assuming sufficient analyticity, one then defines Wick rotation of
3620: the states \eqref{formal extension 2} by the contour deformation of
3621: $\cC$ accompanied by the continuation $q\,\rightarrow\, i\om, x
3622: \rightarrow \, i t$;
3623: \begin{align}
3624: \bra{B'} \Big\vert_{\rm Lorentzian} :=
3625: \sum_{\al}\int_{-\infty}^{\infty} \frac{i \dd \om}{2\pi}\,
3626: \int_{-\infty}^{\infty} i\dd t\, f(it, \al) e^{-i\om t} \,
3627: \dbra{i\om,\al} ~. \label{Wick rotation 0}
3628: \end{align}
3629: This is essentially the procedure taken in \cite{Nakayama:2004yx}. Of course, we
3630: potentially have an ambiguity in the choice of the contour $\cC$,
3631: and the correct choice should be determined by the physics under
3632: study.
3633: 
3634: In the present case $\bra{B}$ corresponds to \eqref{clss2'} and
3635: $\bra{B'}$ is given by
3636: \begin{align}
3637: {}_{\rm D1}\bra{B';\rho_0,\theta_0} = \int_0^{\infty} \frac{\dd
3638: p}{2\pi}\, \int_{-\infty}^{\infty} \frac{\dd q}{2\pi}\, \Psi'_{\rm
3639: D1} (\rho_0,\theta_0;p,q) \, \dbra{p,q} ~, \end{align}
3640: %
3641: where
3642: %
3643: \begin{align} &\Psi'_{\rm D1}(\rho_0,\theta_0;p,q)
3644: \cr
3645:  &= \frac{\sinh(\pi p)}
3646: {\left|\cosh\Big(\pi\frac{p+iq}{2}\Big)\right|^2} \,
3647: \frac{\pi\Gamma(ip)\Gamma\Big(1+\frac{ip}{k}\Big)}
3648: {\Gamma\Big(\frac{1}{2}+\frac{ip+q}{2}\Big)
3649: \Gamma\Big(\frac{1}{2}+\frac{ip-q}{2}\Big)} \, e^{iq\theta_0}
3650: \left[e^{-ip\rho_0} +\frac {\cosh\left(\pi\frac{p-i|q|}{2}\right)}
3651: {\cosh\left(\pi\frac{p+i|q|}{2}\right)}
3652:  e^{ip\rho_0}\right] \cr
3653:  &\equiv  B\left(\frac{1}{2}-\frac{ip-q}{2},
3654: \frac{1}{2}-\frac{ip+q}{2}\right) \Gamma\left(1+\frac{ip}{k}\right)
3655: \, e^{iq\theta_0}\left[e^{-ip\rho_0}
3656: +\frac{\cosh\left(\pi\frac{p-i|q|}{2}\right)}
3657: {\cosh\left(\pi\frac{p+i|q|}{2}\right)} e^{ip\rho_0}\right].
3658: \label{D1'}
3659: \end{align}
3660: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3661: \begin{figure}[htbp]
3662:     \begin{center}
3663: %    \includegraphics[width=0.5\linewidth,keepaspectratio,clip]
3664:    \includegraphics[width=13cm,height=10cm]
3665:       {cont0.eps}
3666:     \end{center}
3667:     \caption{The red (green broken) line
3668: is the contour $\cC^+$ for $p > 0$
3669:      to the Lorentzian time. Notice that an infinite number
3670:      of branch cuts repeats in the Euclidean time: $\frac{\pi}{2}+2n\pi < x <
3671:       \frac{3\pi}{2} +2n\pi$, $(n\in \bz)$ along the real $x$-axis.}
3672:     \label{c-array}
3673: \end{figure}
3674: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3675: Here $B(p,q) \equiv \Gamma(p)\Gamma(q)/\Gamma(p+q)$ denotes Euler's
3676: beta function. The integration contour $\cC$ we choose is shown in
3677: Figure \ref{c-array} \cite{Nakayama:2004yx}. As in \eqref{evaluation overlap phi},
3678: we separately evaluated the integrals of $\phi^{p}_{L,q}$ and
3679: $\phi^p_{R,q}$ based on the decomposition \eqref{decomp ef}. For the
3680: convergence of integrals, we choose the contour $\cC^+$ for
3681: $\phi^{p}_{L,q}$ ($p>0$ sector) and $\cC^-$ for  $\phi^{p}_{R,q}$
3682: ($p<0$ sector). Such choice of integration contours rendered an
3683: extra damping factor $\sinh(\pi p) /
3684: {|\cosh\left(\pi\frac{p+iq}{2}\right)|^2}$, which improves the
3685: ultraviolet behavior of the wavefunction and makes it possible to
3686: take the Wick rotation sensibly. The non-trivial phase factor $
3687: {\cosh\left(\pi\frac{p-i|q|}{2}\right)}/
3688: {\cosh\left(\pi\frac{p+i|q|}{2}\right)} $ in the second term
3689: originates from the reflection amplitude,
3690: %(recall \eqref3{decomp ef}
3691: %$\phi^{p}_n=\phi^{p}_{L,n}+ \cR_0 \phi^p_{R,n}$)
3692: and it reduces to $(-1)^n$ when $q=n\in \bz$.
3693: 
3694: The second subtlety implies that $\dbra{i\om, \al}$ is not uniquely
3695: defined in \eqref{Wick rotation 0}. This is the issue that arises in a
3696: background with horizon, equivalently, non-existence of globally
3697: definable timelike Killing vector. As such, this subtlety did not
3698: arise for the extremal NS5-brane geometry (described asymptotically
3699: by free linear dilaton theory) considered in
3700: \cite{Nakayama:2004yx}. In section \ref{sec:7-2-4}, within the mini-superspace analysis
3701: for the Lorentzian two-dimensional black hole, we shall clarify this
3702: subtlety.
3703: 
3704: An alternative, sensible prescription of the analytic continuation
3705: is to define the disk one-point correlator {\em directly\/} via the
3706: Lorentzian Fourier transform:
3707: \begin{align}
3708: & \Big< \Phi^{\msc{Lorentzian}}_{\om,\al} \Big>_{\msc{disk}}
3709:    = \int_{-\infty}^{\infty} \dd t\, f(it,\al) e^{-i\om t}~.
3710: \label{formal disk amp 2}
3711: \end{align}
3712: This is {\em not\/} always equivalent to the the former method
3713: elaborated above. In fact, the latter method does not necessarily
3714: assert that the boundary state constructed so is expandable in terms
3715: of the Lorentzian Ishibashi states that are analytically continued
3716: from the Euclidean ones.\footnote{Recently, we have succeeded the direct evaluation of the overlap integral. See appendix \ref{direct} for details.}
3717: 
3718: 
3719: In section \ref{sec:3-3-4}, we have reviewed the primary states for the Lorentzian two-dimensional black hole.
3720: Having obtained the Lorentzian primary states, we shall now
3721: construct several interesting class of boundary states for a
3722: D0-brane propagating in the black hole background. We have seen that
3723: the D0-brane propagates along the trajectory \eqref{trajectory D0}.
3724: The two-dimensional black hole is eternal, so, in addition to the
3725: past and the future asymptotic infinities, the causal propagation
3726: region has the past horizon ${\cal H}^-$ surrounding the white hole
3727: singularity and the future horizon ${\cal H}^+$ surrounding the
3728: black hole singularity. As such, by taking variety of possible
3729: boundary conditions, we can construct interesting class of boundary
3730: states.
3731: 
3732: 
3733: 
3734: 
3735: 
3736: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3737: \subsubsection{boundary state of D0-brane absorbed to future horizon}\label{sec:7-2-2}
3738: 
3739: 
3740: 
3741: 
3742: Consider first the boundary state obeying the boundary condition
3743: $\psi(\rho,t)\,\rightarrow\, 0$ at the past horizon ${\cal H}^-$,
3744: viz. the primary states $\ket{U^p_{\om}}$.
3745: %for $\om >0$ and $\ket{V^p_{\om}}$ for $\om <0$.
3746: This boundary condition is relevant for scattering of a D0-brane off
3747: the black hole, since the condition represents absorption only and
3748: no emission of the D0-brane by the black hole. D0-brane boundary
3749: state obeying such absorbing boundary condition is then expanded
3750: solely by the Ishibashi states ${}^{\widehat{U}}\dbra{p,\om}$,
3751: $\dket{p,\om}^U$ that are associated with the primary states
3752: $\widehat{\bra{U^p_{\om}}}$, $\ket{U^p_{\om}}$:
3753: \begin{align}
3754:  &
3755: {}_{\msc{absorb}}\!\bra{B;\rho_0,t_0} = \int_0^{\infty}\frac{\dd
3756: p}{2\pi} \int_{-\infty}^{\infty}\frac{\dd \om}{2\pi}\,
3757:   \Psi_{\msc{absorb}}(\rho_0,t_0;p,\om) \,
3758: {}^{\widehat{U}}\!\dbra{p,\om}~, \nn
3759: %%%
3760:   &
3761: \ket{B;\rho_0,t_0}_{\msc{absorb}} = \int_0^{\infty}\frac{\dd
3762: p}{2\pi} \int_{-\infty}^{\infty}\frac{\dd \om}{2\pi}\,
3763:   \Psi^*_{\msc{absorb}}(\rho_0,t_0;p,\om)
3764: \, \dket{p,\om}^U~. \label{falling D0 0}
3765: \end{align}
3766: The boundary wavefunction $\Psi_{\msc{absorb}}(\rho_0,t_0;p,\om)$
3767: is then interpreted as the disk one-point correlators:
3768: \begin{align}
3769: \Psi_{\msc{absorb}}(\rho_0,t_0;p,\om) &= \langle U^p_{\om}
3770: \rangle_{\msc{disk}} \equiv {}_{\msc{absorb}}\!\bra{B;\rho_0,t_0}
3771: U^p_{\om} \rangle~,
3772: %   (\mbox{if} ~ \om >0) \nn
3773: %&=& \langle V^{-p}_{\om} \rangle_{\msc{disk}}
3774: %\equiv
3775: %{}_{\msc{falling}}\!\bra{B;\rho_0,t_0} V^{-p}_{\om}
3776: %\rangle~, ~~~
3777: %   (\mbox{if} ~ \om <0) ~.
3778: \label{falling disk}
3779: \end{align}
3780: 
3781: 
3782: The boundary wavefunction \eqref{falling disk} is then obtained by
3783: taking the Wick rotation $q\,\rightarrow\, i\om$ ($q\,\rightarrow\,
3784: -i\om$) for $q<0$ ($q>0$)
3785: %$\om>0$ ($\om<0$)
3786: in \eqref{D1'} (recall \eqref{ac UVLR}):\footnote
3787:         {In reality, there is a further overall factor $i$,
3788:           but, for notational simplicity, we will absorb it to the definition
3789:           of the Ishibashi states.}
3790: \begin{align}
3791: & \hspace{-1cm} \Psi_{\msc{absorb}}(\rho_0,t_0;p,\om) = B(\nu_+,
3792: \nu_-) \Gamma\Big(1+\frac{ip}{k}\Big) \, e^{-i\om t_0}\left[ e^{-ip
3793: \rho_0} -  \frac{\cosh\left(\pi \frac{p-\om}{2}\right)}
3794: {\cosh\left(\pi \frac{p+\om}{2}\right)} e^{ip\rho_0 } \right]~,
3795: \label{falling D0}
3796: \end{align}
3797: The relative minus sign in the second term of
3798: $\Psi_{\msc{absorb}}(\rho_0,t_0;p,\om)$ originates from the fact
3799: that the contour rotation defining the Wick rotation has opposite
3800: directions for $\cC^+$ (suitable for $p>0$) and $\cC^-$ (suitable
3801: for $p<0$). See figure \ref{c-array}. This boundary wavefunction
3802: \eqref{falling D0} satisfies the exact reflection relation
3803: \begin{align}
3804: \Psi_{\msc{absorb}}(\rho_0,t_0;-p,\om)= \cR(-p,\om) \,
3805: \Psi_{\msc{absorb}}(\rho_0,t_0;p,\om)~. \label{ref falling D0}
3806: \end{align}
3807: 
3808: With such boundary condition, the boundary wavefunction
3809: \eqref{falling D0} would have no overlap with D0-brane's trajectory
3810: \eqref{trajectory D0} in the far past region $t\ll t_0$. In fact, the
3811: trajectory \eqref{trajectory D0} starts from the past horizon ${\cal
3812: H}^-$ at $t=-\infty$, reaches the time-symmetric point $\rho =
3813: \rho_0$ at $t = t_0$, and then falls back the future horizon ${\cal
3814: H}^+$ at $t=+\infty$, while the wavefunction $U^p_{\om}$ does not
3815: have any component outgoing from ${\cal H}^-$. We thus interpret
3816: that the boundary state \eqref{falling D0} describes the future half
3817: of the classical trajectory \eqref{trajectory D0}. We shall hence call
3818: it the `absorbed D-brane'.
3819: 
3820: By utilizing the radion-tachyon correspondence, the rolling radion
3821: (as described by the boundary state \eqref{falling D0}) can be also
3822:   interpreted as the rolling tachyon. In the latter interpretation,
3823: the D0-brane absorbed to the future horizon is the counterpart of
3824: the future-half S-brane \cite{Gutperle:2002ai,Strominger:2002pc,Larsen:2002wc}, in which the
3825: tachyon rolls down the potential hill at asymptotic future $t
3826: \rightarrow + \infty$ and emits radiation.
3827: 
3828: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3829: \subsubsection{boundary state of D0-brane emitted from past horizon}\label{sec:7-2-3}
3830: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3831: Consider next the boundary condition: $\psi(\rho,t)\,\rightarrow\,
3832: 0$ at ${\cal H}^+$, viz. use the basis $\dket{p,\om}^V$,
3833: ${}^{\widehat{V}}\!\dbra{p,\om}$ instead of $\dket{p,\om}^U$,
3834: ${}^{\widehat{U}}\!\dbra{p,\om}$. Utilizing the reflection relation,
3835: we can first rewrite \eqref{D1'} as the form which only includes the
3836: $p<0$ Ishibashi states by means of the reflection relation. Then, we
3837: can analytically continue the states $\ket{\phi^{-p}_q}$ ($p>0$)
3838: into $\ket{V^p_\om}$.
3839: %(associated to the $\om>0$ part of the basis $\dket{p,\om}^-$).
3840: The resultant boundary state is obtained by
3841: simply replacing $p\,\rightarrow\,-p$,
3842: $\om\, \rightarrow\, -\om$
3843: in \eqref{falling D0};
3844: \begin{align}
3845: & {}_{\msc{emitted}}\!\bra{B;\rho_0,t_0} = \int_0^{\infty}\frac{\dd
3846: p}{2\pi} \int_{-\infty}^{\infty}\frac{\dd \om}{2\pi}\,
3847:   \Psi_{\msc{emitted}}(\rho_0,t_0;p,\om) \,
3848: {}^{\widehat{V}}\!\dbra{p,\om}~. \nn
3849: %%%
3850: & \ket{B;\rho_0,t_0}_{\msc{emitted}} = \int_0^{\infty}\frac{\dd
3851: p}{2\pi} \int_{-\infty}^{\infty}\frac{\dd \om}{2\pi}\,
3852:   \Psi^*_{\msc{emitted}}(\rho_0,t_0;p,\om) \,
3853: \dket{p,\om}^V ~. \label{emitted D0} \end{align}
3854: %
3855: where
3856: %
3857: \begin{align} \Psi_{\msc{emitted}}(\rho_0,t_0;p,\om) = B(\nu^*_+, \nu^*_-)
3858: \Gamma\left(1-\frac{ip}{k}\right) \, e^{-i\om t_0}\left[
3859: e^{ip\rho_0} - \frac{\cosh\left(\pi \frac{p-\om}{2}\right)}
3860: {\cosh\left(\pi \frac{p+\om}{2}\right)} e^{-ip\rho_0} \right]~.
3861: \nonumber
3862: \end{align}
3863: Obviously, the emitted D0-brane wavefunction is the time-reversal
3864: of the absorbed D0-brane wavefunction \eqref{falling D0}:
3865: %
3866: \begin{align} \Psi_{\msc{emitted}}(\rho_0,t_0;p,\om) =
3867: \Psi^*_{\msc{absorb}}(\rho_0,-t_0;p,\om) ~. \nonumber \end{align}
3868: %
3869: Namely, it describes the D0-brane emitted from the past horizon at
3870: asymptotic past $t=-\infty$. By the choice of the boundary
3871: condition, this boundary state \eqref{emitted D0} describes only the
3872: past half of the classical D0-brane trajectory \eqref{trajectory D0}.
3873: 
3874: The exact reflection relation has the form
3875: \begin{align}
3876: & \Psi_{\msc{emitted}}(\rho_0,t_0;-p,\om)= \cR^*(-p,\om) \,
3877: \Psi_{\msc{emitted}}(\rho_0,t_0;p,\om)~. \label{ref emiited D0}
3878: \end{align}
3879: 
3880: Again, in light of the radion-tachyon correspondence, the D0-brane
3881: emitted from the past horizon is the counterpart of the past-half
3882: S-brane in tachyon rolling. The radiation creeps up the tachyon
3883: potential hill from past infinity and forms an unstable D-brane.
3884: 
3885: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3886: \subsubsection{boundary state of time-symmetric D0-brane}\label{sec:7-2-4}
3887: 
3888: The third possible boundary state is obtainable by {\em directly\/}
3889: taking the analytic continuation in the disk one-point amplitudes,
3890: as we already mentioned. Recalling \eqref{ac UVLR}, we shall
3891: analytically continue the disk amplitudes as (assume $p>0$)
3892: \begin{align}
3893:   \langle \phi^{+p}_q \rangle_{\msc{disk}}~\longrightarrow~
3894:   \langle U^p_{\om} \rangle_{\msc{disk}} \qquad \mbox{and} \qquad
3895:   \langle \phi^{-p}_q \rangle_{\msc{disk}}~\longrightarrow~
3896:   \langle V^p_{\om} \rangle_{\msc{disk}}~.
3897: \label{ac disk amp}
3898: \end{align}
3899: The Euclidean one-point amplitudes $\langle \phi^{\pm p}_q
3900: \rangle_{\msc{disk}}$ are given in \eqref{D1'}, and can be expressed
3901: in contour integrals as in \eqref{formal disk amp 1}. Recall that
3902: $\langle \phi^p_{L,q}   \rangle_{\msc{disk}}$, $\langle
3903: \phi^{p}_{R,q} \rangle_{\msc{disk}}$ are prescribed by the contour
3904: integrals over $\cC^+$, $\cC^-$ in figure \ref{c-array}. We shall
3905: thus analytically continue them to the real time axis (imaginary
3906: $x$-axis). In this way, we extract the Lorentzian disk one-point
3907: amplitudes as
3908: \begin{align}
3909: & \langle U^p_{\om} \rangle_{\msc{disk}} = \langle U^p_{\om}
3910: \rangle_{\msc{disk}}^{(\msc{absorb})} \qquad \mbox{and} \qquad \langle V^p_{\om}
3911: \rangle_{\msc{disk}} = \langle V^p_{\om} \rangle_{\msc{disc}}^{(\msc{emitted})}~,
3912: ~~~ \label{rel disc amp}
3913: \end{align}
3914: where the right-hand sides are simply the amplitudes associated with
3915: the `absorbed' and `emitted' D0-branes considered in the previous
3916: subsections and explicitly given in \eqref{falling D0} and
3917: \eqref{emitted D0}. Since $U^p_{\om}$ and $V^p_{\om}$ constitute the
3918: complete set of basis for Lorentzian primary fields, the amplitudes
3919: \eqref{rel disc amp} would yield yet another Lorentzian D0-brane
3920: boundary states. As is obvious from the above construction, this
3921: state keeps the time-reversal symmetry manifest and reproduces the
3922: entire classical trajectory \eqref{trajectory D0}, that is, it
3923: describes a D0-brane emitted from the past horizon and reabsorbed to
3924: the future horizon. From the viewpoint of the boundary conformal
3925: theory, this would be considered the most natural one since it
3926: captures the entire classical trajectory of the D0-brane. In the
3927: radion-tachyon correspondence, this state is the counterpart of the
3928: full S-brane \cite{Gutperle:2002ai,Sen:2002nu,Sen:2002in,Sen:2002an}.
3929: 
3930: Explicitly, the time-symmetric boundary states are given by
3931: %%%%%
3932: \begin{align}
3933: &{}_{\msc{symm}}\!\bra{B;\rho_0,t_0} \cr
3934: &={}_{\msc{absorb}}\!\bra{B;\rho_0,t_0} +
3935: {}_{\msc{emitted}}\!\bra{B;\rho_0,t_0} \cr &=
3936: \int_0^{\infty}\frac{\dd p}{2\pi} \int_{-\infty}^{\infty}\frac{\dd
3937: \om}{2\pi}\, \left[
3938:   2\Psi_{\msc{symm}}(\rho_0,t_0;p,\om)
3939: \, {}^L\!\dbra{p,\om}
3940:   +
3941:   2\Psi^*_{\msc{symm}}(\rho_0,-t_0;p,\om) \, {}^R\!\dbra{p,\om}
3942: \right]  \cr
3943: %%% 
3944: & \ket{B;\rho_0,t_0}_{\msc{symm}} \cr &=
3945: \ket{B;\rho_0,t_0}_{\msc{absorb}} +
3946: \ket{B;\rho_0,t_0}_{\msc{emitted}}  \cr  &=
3947: \int_0^{\infty}\frac{\dd p}{2\pi} \int_{-\infty}^{\infty}\frac{\dd
3948: \om}{2\pi}\, \left[
3949:   2\Psi^*_{\msc{symm}}(\rho_0,t_0;p,\om)
3950: \, \dket{p,\om}^L
3951:   +
3952:   2\Psi_{\msc{symm}}(\rho_0,-t_0;p,\om) \, \dket{p,\om}^R
3953: \right] ~, \label{symmetric D0} 
3954: \end{align}
3955: %
3956: where
3957: %
3958: \begin{align} \Psi_{\msc{symm}}(\rho_0,t_0;p,\om) = B(\nu_+, \nu_-)
3959: \Gamma\left(1+\frac{ip}{k}\right) \, e^{-ip\rho_0-i\om t_0}
3960: \end{align}
3961: %
3962: and ${}^L\!\dbra{p,\om}$, $\dket{p,\om}^L$, ${}^R\!\dbra{p,\om}$,
3963: $\dket{p,\om}^R$ are the Ishibashi states constructed over the
3964: primary states $\bra{L^p_{\om}}$, $\ket{L^p_{\om}}$,
3965: $\bra{R^p_{\om}}$, $\ket{R^p_{\om}}$,\footnote{The extra factor of
3966: `2' was introduced for convenience. Recall \eqref{inner product
3967: UVLR}.} respectively. One can readily check that the second lines
3968: in \eqref{symmetric D0} are indeed correct by evaluating the disk
3969: one-point amplitudes from them. For instance, using \eqref{inner
3970: product UVLR}, we obtain
3971: \begin{align}
3972:  \langle U^p_{\om} \rangle_{\msc{disk}}^{(\msc{symm})}
3973: &= {}_{\msc{symm}}\!\bra{B;\rho_0,t_0} U^p_{\om} \rangle \nn &=
3974: \Psi_{\msc{symm}}(\rho_0,t_0;p,\om) +\cR(p,\om)
3975: \Psi^*_{\msc{symm}}(\rho_0,-t_0;p,\om) \nn &=  B(\nu_+,\nu_-)
3976: \Gamma\left(1+\frac{ip}{k}\right) e^{-i\om t_0} \, \left[
3977: e^{-ip\rho_0} - \frac{\cosh\left(\pi \frac{p-\om}{2}\right)}
3978: {\cosh\left(\pi \frac{p+\om}{2}\right)} e^{ip\rho_0} \right] \nn &=
3979: \langle U^p_{\om} \rangle^{(\msc{absorb})}_{\msc{disk}} \equiv
3980: {}_{\msc{absorb}}\!\bra{B;\rho_0,t_0}U^p_{\om}\rangle ~.
3981: \end{align}
3982: Other one-point amplitudes can be checked analogously.
3983: 
3984: Two remarks are in order. First, notice that, though the disk
3985: one-point amplitudes are, the symmetric boundary states
3986: \eqref{symmetric D0} by themselves are {\em not\/} analytically
3987: continuable to the Euclidean boundary state \eqref{D1'}. This should
3988: not be surprising as the Lorentzian Hilbert space is generated by
3989: {\sl twice} as many generators as the Euclidean theory.
3990: In other words,
3991: %$\ket{L^p_{\om}}$, $\ket{R^p_{\om}}$ do not have normalizable
3992: %counterparts in the Euclidean Hilbert space.
3993: %%%%
3994: the Lorentzian bases $\ket{U^p_{\om}}$, $\ket{V^p_{\om}}$ correspond
3995: to $\ket{\phi^p_n}$, $\ket{\phi^{-p}_n}$ in the Euclidean theory,
3996: which were however linearly dependent due to the reflection
3997: relation.
3998: %%%%
3999: Nevertheless, the
4000: boundary state \eqref{symmetric D0} is a consistent one and yields
4001: disk one-point amplitudes that can be correctly continued to the
4002: Euclidean ones. Second, the full Lorentzian Hilbert space is
4003: decomposed as
4004: \begin{align}
4005:  \cH = \cH^U \oplus \cH^V  \qquad \mbox{and} \qquad
4006:  \widehat{\cH} = \widehat{\cH^U} \oplus \widehat{\cH^V}~,
4007: \label{decomp Hilb}
4008: \end{align}
4009: where $\cH^U$ ($\cH^V$) is spanned by $\ket{U^p_{\om}}$ , ($\,
4010: \ket{V^p_{\om}}\, $) and their descendants. The dual space
4011: $\widehat{\cH^U}$ ($\widehat{\cH^V}$) is similarly spanned by
4012: $\widehat{\bra{U^p_{\om}}}$, ($\widehat{\bra{V^p_{\om}}}$). Here,
4013: the Hilbert subspaces $\cH^{U}$, $\widehat{\cH^{U}}$ ($\cH^{V}$,
4014: $\widehat{\cH^{V}}$) correspond to the boundary condition
4015: $\psi(\rho,t)\,\rightarrow\, 0$ at ${\cal H}^{-}$ (${\cal H}^+$).
4016: The `absorbed' and `emitted' D0-brane boundary states \eqref{falling
4017: D0}, \eqref{emitted D0} are consistent {\sl only} in the subspaces
4018: $\cH^U$, $\cH^V$ ($\widehat{\cH^U}$, $\widehat{\cH^V}$), while the
4019: `symmetric' D0-brane boundary state \eqref{symmetric D0} is
4020: well-defined in the entire Hilbert space $\cH$ ($\widehat{\cH}$). We
4021: thus have simple relations
4022: \begin{align}
4023: \ket{B;\rho_0,t_0}_{\msc{absorb}} &= P_U \,
4024: \ket{B;\rho_0,t_0}_{\msc{symm}} & \mbox{and} \qquad 
4025: {}_{\msc{absorb}}\! \bra{B;\rho_0,t_0} &=
4026: {}_{\msc{symm}}\!\bra{B;\rho_0,t_0}\,\widehat{P_U}~,\cr
4027: \ket{B;\rho_0,t_0}_{\msc{emitted}} &= P_V \,
4028: \ket{B;\rho_0,t_0}_{\msc{symm}} &  \mbox{and} \qquad
4029:  {}_{\msc{emitted}}\!\bra{B;\rho_0,t_0} &=
4030: {}_{\msc{symm}}\!\bra{B;\rho_0,t_0}\,\widehat{P_V}~,
4031: \label{proj symmetric D0}
4032: \end{align}
4033: where $P_{U, V}$ ($\widehat{P_{U,V}}$) denotes projection of the
4034: Hilbert space ${\cal H}$ to $\cH^{U,V}$ ($\widehat{\cH^{U, V}}$).
4035: 
4036: 
4037: 
4038: 
4039: 
4040: 
4041: 
4042: \subsection{Rolling D-brane gathers moss}\label{sec:7-3}
4043: As we have discussed in section \ref{sec:4}, it is of critical importance to study the $1/k$ corrections to the boundary states in order to understand the ``black hole - string transition" probed by our rolling D-brane. We first note that the boundary wavefunction itself is an analytic function with respect to $k$,\footnote{A possible exception is the factor $\nu_b^{2j+1}$. However, this factor can be absorbed (renormalized) into the cosmological constant operator of the $\mathcal{N}=2$ Liouville theory, or the mass of the two-dimensional black hole, so we will neglect this small subtlety.} so the boundary wavefunction itself is a well-defined quantity even for $k<1$. An alternative way to confirm this is to note that, at least in the Euclidean signature, the boundary wavefunction satisfies the conformal bootstrap equation for the dual $\mathcal{N}=2$ Liouville theory whose description is more reliable for $k<1$.
4044: 
4045: To see the effect of the $1/k$ corrections clearly, it is convenient to go to the coordinate space representation rather than the momentum space representation. For simplicity, we take the linear dilaton (extremal) limit of the boundary wavefunction (see section \ref{sec:8-5} for details of this limit). In the momentum space we have,
4046: \begin{align}
4047:  \Psi(\rho_0,t_0;p,\om) = \frac{1}{2} B(\nu_+,\nu_-)
4048: \Gamma\left(1+i\frac{p}{k}\right)\, e^{-ip\rho_0-i\om t_0}~,
4049:  \quad \mbox{where} \quad \nu_{\pm}
4050: \equiv \frac{1}{2}- i\frac{p\pm \om}{2}~.
4051: \end{align}
4052: 
4053: 
4054: We can Fourier transform this boundary wavefunction to obtain the boundary wavefunction in the coordinate space:
4055: \begin{align}
4056: \Psi(\rho,t) = \frac{\sqrt{k}}{\pi(2\cosh t)^{k+1}} \exp\left[-k\rho- \frac{e^{-k\rho}}{(2\cosh t)^{k}} \right] \ .
4057: \end{align}
4058: It will be localized along the classical trajectory:
4059: \begin{align}
4060: \rho_0(t) = - \log(2\cosh t) \ 
4061: \end{align}
4062: in the semiclassical limit (i.e. $k\to \infty$). For finite $k$, the classical trajectory is smeared. 
4063: 
4064: To go further, we study the energy momentum distribution for finite $k$. Expanding the boundary states and reading the coupling to the gravity, we obtain (see \cite{Nakayama:2004ge} for details of the computation)
4065: \begin{align}
4066: T_{00} = \left(\frac{e^{-\rho}}{2\cosh t}\right)^{k-1} \exp\left[-\left(\frac{e^{-\rho}}{2\cosh t}\right)^{k}\right] \ .
4067: \end{align}
4068: The distribution of the energy is Poisson type and the maximum of the energy density is now located at
4069: \begin{align}
4070: \frac{e^{-\rho}}{2\cosh t} = 1-\frac{1}{k} \ .
4071: \end{align}
4072: The variance of the distribution is computed as
4073: \begin{align}
4074: \Delta\rho \simeq \sqrt{\frac{1}{2(k-1)}} \ , \label{ditvar}
4075: \end{align}
4076: which can be regarded as the smearing factor for the classical trajectory due to the $\alpha'$ corrections. One might say that the rolling D-brane gathers moss in the $\alpha'$ corrected black hole background. The moss could be identified with the analytic continuation of the winding tachyon \cite{Kutasov:2005rr}. Indeed the similar smearing factor in the Euclidean hairpin brane can be understood from the open string winding tachyon condensation near the tip of the Euclidean hairpin.
4077: 
4078: 
4079: We again emphasize that the coordinate space wavefunction itself is an analytic function with respective $k$. However, below $k=1$, the variance of the smeared D-brane trajectory \eqref{ditvar} diverges, which means that the boundary wavefunction does not have a sensible interpretation as a rolling D-brane in the classical two-dimensional black hole any more. The transition point exactly coincides with the ``black hole - string transition" point we discussed in section \ref{sec:4}. The classical black hole appears no more black hole at this point, and the D-brane cannot role down into the hole as a probe. In section \ref{sec:8}, we compute the closed string radiation rate from the rolling D-brane, and see explicitly that the radiation rate also reveals such a phase transition as expected. As a consequence, we will see that the ``tachyon - radion correspondence" and the universality of the decaying D-brane breaks down.
4080: 
4081: 
4082: As a generalization of the construction, we can introduce the fundamental string charge (electric flux) along the rolling D-brane boundary states. The construction is based on the Lorenz boost technique reviewed in section \ref{sec:5-2-6}. The corresponding boundary states have been studied in \cite{Nakayama:2004ge,Chen:2004vw}.
4083: \newpage
4084: \sectiono{Black Hole - String Transition from Probe Rolling D-brane}\label{sec:8}In this section we compute the closed string radiation rate from the rolling D-brane. The organization of this section is as follows. In section \ref{sec:8-1}, we compute the closed string radiation rate from the closed string perspective. In section \ref{sec:8-2}, we study the same closed string radiation rate from the open string perspective, establishing the consistency between unitarity and channel duality. In section \ref{sec:8-3}, we discuss the black hole - string transition from the probe rolling D-brane. In section \ref{sec:8-4}, the boundary states and radiation in R-R sector are discussed. In section \ref{sec:8-5}, we study the extremal NS5-brane limit. Finally in section \ref{sec:8-6}, we present the physical interpretations of Hartle-Hawking states for rolling D-branes.\footnote{This section is based on \cite{Nakayama:2005pk,Nakayama:2006qm}.}
4085: 
4086: \subsection{Radiation out of rolling D-brane from closed string viewpoint}\label{sec:8-1}
4087: In the background of the black hole, the D0-brane moves along the
4088: geodesic and we have constructed a variety of boundary states
4089: describing the geodesic motion, specified by appropriate boundary
4090: conditions.
4091: 
4092: Both by gravity and by strong string coupling gradient, the
4093: D$p$-brane is pulled in and finds its minimum energy and mass at the
4094: location of the NS5-brane. The D$p$-brane is supersymmetric in flat
4095: space-time, but preserves no supersymmetry in black NS5-brane
4096: background. Even in extremal NS5-brane background, until the
4097: D$p$-brane dissociates into the NS5-brane and form a non-threshold
4098: bound-state, the space-time supersymmetry is completely broken. In
4099: these respects,the D$p$-brane propagating in the NS5-brane
4100: background is much like excited D$p$-brane (many excited open
4101: strings attached on it) in flat space-time. Decay of the latter via
4102: closed string emission was studied extensively for $p=1$
4103: \cite{Callan:1996dv,Das:1996wn}: the decay spectrum was found to match exactly
4104: with the Hawking radiation of the non-extremal black hole made out
4105: of these excited D-branes, and the effective temperature of excited
4106: open string modes agrees exactly with the Hawking temperature. In
4107: this section, we shall find certain analogous results for the closed
4108: string radiation off the rolling D0-brane, though special features
4109: also arise.
4110: 
4111: As the D0-brane is pulled in, acceleration would grow and radiates
4112: off the binding energy into closed string modes. Details of the
4113: radiation spectra would differ for different choice of the boundary
4114: conditions, viz. for different boundary states of the D0-brane. In
4115: this section, as a probe of the black hole geometry and D-brane
4116: dynamics therein, we shall analyze spectral distribution of the
4117: closed string radiation off the rolling D0-particle.
4118: 
4119: By applying the optical theorem, the radiation rate during the
4120: radion-rolling process is obtainable as the imaginary part of the
4121: annulus amplitude in the closed string channel.\footnote{For the
4122: tachyon rolling process in flat space-time background, the amplitude
4123: was evaluated first in \cite{Lambert:2003zr,Karczmarek:2003xm}.} Denote the differential
4124: number density $\dd {\cal N}(p, M)$ of the radiation at a fixed
4125: value of the radial momentum $p$ and the mass-level $M$. By the
4126: definition of the D-brane boundary state, the radiation number
4127: density $\dd \cN$ is then given in terms of the boundary wave
4128: functions:
4129: \begin{align}
4130: \dd \cN (p,M)   &:= {\dd p \over 2 \pi} {\dd M \over (2 \pi)^d}
4131: \int{\dd \om} \, \Big< \Psi(\om, p, M) \Big\vert \delta(L_0 +
4132: \overline{L}_0) \Big\vert \Psi (\om, p, M) \Big>
4133: \nonumber \\
4134: &= \frac{\dd p}{2 \pi} \frac{\dd M}{(2 \pi)^d} \frac{1}{2
4135: \om(p,M)}\, \Big| \Psi(p,\om(p,M))\Big|^2 ~. \label{radiation rate
4136: 0}
4137: \end{align}
4138: Here, $\om, p$ are the energy and the radial momentum in
4139: two-dimensional Lorentzian background, $M$ is the total mass
4140: (conformal weight) of the remaining subspaces of dimension $d$
4141: (including mass gap), $\Psi(\om, p, M)$ is the boundary wave
4142: function (including that of the remaining subspace), and
4143: $\om(p,M)(>0)$ is the on-shell energy of the radiated closed string
4144: state determined by the on-shell condition $L_0 + \overline{L}_0 =
4145: 0$ including the ghost contribution. From the kinematical
4146: consideration, it is obvious that the differential number density
4147: \eqref{radiation rate 0} is nonzero only when the D-brane is rolling.
4148: Of particular physical interest is the spectral distribution in the
4149: phase-space, as measured by the independent moments, {\em e.g.}
4150: %
4151: \begin{align} \Big< \om^m M^n \Big> &= \int \frac{\dd p}{2 \pi} \frac{\dd
4152: M}{(2 \pi)^d} \om^m(p, M) M^n \frac{1}{2 \om(p, M)}  \Big|\Psi(p,
4153: \om(p,M))\Big|^2 \nonumber \end{align}
4154: %
4155: for $m, n =0, 1, 2, \cdots$. We shall evaluate these spectral
4156: observables by first evaluating the integral over the radial
4157: momentum $p$ by saddle-point approximation. In doing so, we pay
4158: particular attention to the asymptotic behavior as the mass-level
4159: $M$ becomes asymptotically large. We shall then evaluate the
4160: integral over the mass-level (conformal weight) $M$, and extract the
4161: spectral observables.
4162: 
4163: %The eternal black hole geometry has two sets of null boundaries:
4164: %past/future horizon and past/future null infinity.
4165: Consider the
4166: boundary state \eqref{falling D0} describing a D0-brane absorbed by
4167: the future horizon. The radiation emitted by the D0-brane is
4168: decomposable into `incoming' (toward the horizon) and `outgoing'
4169: (toward the null infinity) components in the far future. The
4170: positive energy sector is expanded by the wavefunction $U^p_{\om}$,
4171: and has the following asymptotic behavior at $t\rightarrow +\infty$:
4172: \begin{align}
4173:  & U^p_{\om}(\rho,t) \sim e^{- i\om \ln \rho -i \om t}
4174:   + d(p, \om) e^{-\rho} e^{+ ip \rho -i\om t} \qquad \mbox{where}
4175:   \qquad  |d(p, \om)| \sim e^{-\pi p}~.
4176: \label{as U}
4177: \end{align}
4178: Here, we assumed $\om \sim M \gg 0$. The first and the second terms
4179: correspond to the incoming wave supported around $\rho=0$ and the
4180: outgoing wave supported in the region $\rho\sim +\infty$,
4181: respectively. The damping factor $d(p)$ originates from the exact
4182: reflection amplitude $\cR(p,\om)$. (See \eqref{decomp ef 2},
4183: \eqref{decomp ef 2-2}.) To obtain the radiation number density, we
4184: need to evaluate $\left|\Psi(p,\om)\right|^2 \times
4185: |U^{p}_{\om}(\rho,t)|^2$. At far future infinity, the interference
4186: term in $|U^p_\om|^2$ drops off upon taking the $p$-integral.
4187: Therefore, after integrating over the radial momentum $p$, the
4188: partial radiation distribution is seen to consist of the `incoming'
4189: and `outgoing' parts:
4190: \begin{align}
4191:  \cN(M)_{\msc{in}} &\equiv \int_0^M \dd M {\dd \cN_{\msc{in}}
4192:  \over \dd M} = \int_0^{\infty} \frac{\dd p}{2 \pi} \frac{1}{2\om(p,M)}
4193:  \Big|\Psi(p,\om(p,M))\Big|^2 \nn
4194: \cN(M)_{\msc{out}} &\equiv \int_0^M \dd M {\dd \cN_{\msc{out}}
4195: \over \dd M} = \int_0^{\infty} \frac{\dd p}{2
4196: \pi}\frac{1}{2\om(p,M)} \Big|d(p) \Big|^2
4197: \Big|\Psi(p,\om(p,M))\Big|^2~. \label{radiation rate 1}
4198: \end{align}
4199: We shall now evaluate the branching ratio between the two radiation
4200: rates \eqref{radiation rate 1} with emphasis on possible string
4201: world-sheet effects. To this end, consider the conformal field theory
4202: defined by $SL(2;\br)/U(1) \times \cM$, where $SL(2;\br)/U(1)$
4203: denotes the (super)coset model and $\cM$ denotes a unitary
4204: (super)conformal field theory of central charge $c_\cM$. Such
4205: (super)conformal field theory covers a variety of interesting string
4206: theory backgrounds. For the fermionic string, superconformal
4207: invariance asserts that the central charge ought to be critical:
4208: \begin{align}
4209:  3\Big(1 + \frac{2}{k}\Big) + c_\cM =15~, \nonumber
4210: \end{align}
4211: where $k$ denotes the level of the super $SL(2;\br)$ current
4212: algebra. If the background describes a stack of black NS5-branes,
4213: $\cM= SU(2)_{k} \times \br^5$ where $k$ equals to the NS5-brane
4214: charge. Likewise, for the bosonic string case, conformal invariance
4215: asserts that the central charge should take the critical value:
4216: \begin{align}
4217:  2 + \frac{6}{\kappa-2} + c_{\cM} =26~,
4218: %~~~ \cQ=\sqrt{\frac{2}{\kappa-2}}~,
4219: \end{align}
4220: where now $\kappa$ refers to the level of the bosonic $SL(2;\br)$
4221: current algebra. For the background describing the black hole in
4222: two-dimensional string theory, $\cM$ is empty and $\kappa$ should be
4223: set to $9/4$.
4224: 
4225: It would be illuminating to analyze the branching ratio for the `rolling closed
4226: string', viz. a closed string state of fixed transverse mass $M$ and
4227: radial momentum $p$ propagating in black hole geometry. The
4228: branching ratio is simply given by the reflection amplitude (see
4229: \eqref{exactra}):
4230: %
4231: \begin{align} \left. {{\cal N}_{\rm out}(p, \om) \over {\cal N}_{\rm in } (p,
4232: \omega)} \right|_{\rm closed \, string} = |{\cal R}(p, \om)|^2 =
4233: {\cosh^2 \pi \left(\om-p \over 2 \right) \over \cosh^2 \pi \left(
4234: {\om+p \over 2}\right)}~. \label{tachyon} \end{align}
4235: %
4236: As emphasized below \eqref{exactra}, string world-sheet effects are
4237: present for the reflection amplitude ${\cal R}$ itself but, being an
4238: overall phase, it drops out of \eqref{tachyon}. The $k$-dependence
4239: enters in the branching ratio \eqref{tachyon} only through the
4240: on-shell dispersion relation $\omega = \sqrt{p^2 + 2 k M^2}$. For
4241: two-dimensional case, first studied in \cite{Dijkgraaf:1992ba} and \cite{Giveon:2003wn},
4242: $k=1/2$, $M=0$ and $\omega = p$, so the scattering probability is
4243: exponentially suppressed as the energy increases.
4244: 
4245: For a fixed transverse mass $M$ {\sl and} the forward radial
4246: momentum $p$, the reflection probability of the infalling D0-brane
4247: is given precisely by the same result as \eqref{tachyon}:
4248: %
4249: \begin{align} \left. {{\cal N}_{\rm out}(p, \om) \over {\cal N}_{\rm in } (p,
4250: \omega)} \right|_{\rm D0-brane} = |{\cal R}(p, \om)|^2 = {\cosh^2
4251: \pi \left(\om-p \over 2 \right) \over \cosh^2 \pi \left( {\om+p
4252: \over 2}\right)}~. \end{align}
4253: %
4254: This is simply because back-scattering of the boundary wave function
4255: originates from that of the closed string wave function: roughly
4256: speaking, the boundary wave function is defined by overlap of the
4257: closed string wave function with the classical trajectory of the
4258: D0-brane.
4259: 
4260: Radiation out of the falling D0-brane is coherent, so we integrate
4261: over the radial momentum $p$ as in \eqref{radiation rate 1} in
4262: extracting the branching ratio. We shall first analyze the partial
4263: radiation distribution at large mass-level, $M \rightarrow \infty$.
4264: More precisely, we shall examine asymptotic behavior of ${\cal N}
4265: (M)$ multiplied by the phase-space `degeneracy factor' $\rho(M)\sim
4266: e^{\frac{1}{2}M \beta_{\rm Hg}}$, where $\beta_{\rm Hg}$ denotes
4267: inverse of the Hagedorn temperature. The closed string states that
4268: couple to the boundary states are left-right symmetric, so we need
4269: to take the square root of the usual degeneracy factor in the closed
4270: string sector. Here, inverse of the Hagedorn temperature is given by
4271: \begin{align}
4272: \beta_{\rm Hg} = 4\pi \sqrt{1-\frac{1}{2k}}~, ~~~ \label{Hagedorn
4273: super}
4274: \end{align}
4275: for the superstring theory, and
4276: \begin{align}
4277: \beta_{\rm Hg} = 4\pi \sqrt{2-\frac{1}{2(\kappa-2)}}~, ~~~
4278: \label{Hagedorn bosonic}
4279: \end{align}
4280: for the bosonic string theory, where the $1/k $
4281: $(1/\kappa)$-correction is interpreted as the string world-sheet
4282: effects of the two-dimensional background. These results are
4283: derivable from the Cardy formula with the `effective central charge'
4284: $c_{\msc{eff}}=c-24 h_{\msc{min}}
4285: %\equiv c - \frac{6}{k} ~ (c- \frac{6}{\kappa-2})
4286: $ \cite{Kutasov:1990ua}, where $h_{\msc{min}}$ refers to the lowest conformal
4287: weight of normalizable primary states.
4288: 
4289: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4290: \subsubsection{radiation distribution in superstring theory}\label{sec:8-1-1}
4291: \label{radiation super}
4292: %
4293: Let us begin with the spectral distribution in
4294: superstring theories. We shall focus exclusively on the NS-NS sector
4295: of the radiation and defer the analysis of the R-R sector to section
4296:  \ref{sec:8-4}. The on-shell condition of closed string state in NS-NS sector is
4297: given by
4298: %\footnote
4299: %  {Here we take the convention
4300: %  $j= -\frac{1}{2}+ i\frac{p}{\cQ}$, and
4301: %   $J^3_0=i\omega$.}
4302: \begin{align}
4303: & -\frac{\om^2}{4k} + \frac{p^2}{4k}+ \frac{1}{4k}+ \Delta_\cM =
4304: \frac{1}{2}~, \label{on-shell super}
4305: \end{align}
4306: %
4307: where $\Delta_\cM$ denotes the conformal weight of the $\cM$-part.
4308: The on-shell energy is given by
4309: %
4310: \begin{align}
4311: \om \equiv \om(p,M) = \sqrt{p^2+2k M^2} \qquad \mbox{where} \qquad
4312: M^2 \equiv 2 \left(\Delta_\cM + \frac{1}{4k} - \frac{1}{2} \right)~.
4313: \nonumber \end{align}
4314: %
4315: Consider now a D0-brane propagating outside the black hole and
4316: absorbed into the future horizon. The relevant boundary wave
4317: function was constructed in \eqref{falling D0} and, from them, the
4318: differential radiation number distributions \eqref{radiation rate 1}
4319: can be computed. At large $\omega$ and $p$, using Stirling's
4320: approximation, we find that
4321: \begin{align}
4322: \cN (M)_{\msc{in}} &= \int_0^{\infty}\frac{\dd p}{2 \pi} \frac{1}
4323: {2\om(p,M)}\, \Big| \Psi_{\rm absorb} (\rho_0,t_0;p,\om(p,M))\Big|^2
4324: \nonumber
4325: \\
4326: &\sim {1 \over M} \int_0^{\infty}{\dd p}\,
4327: e^{+\pi\left(1-\frac{1}{k}\right)p - \pi\sqrt{p^2+2k M^2}}~ \label{N
4328: M super in}\\
4329: \nonumber \\
4330: \cN (M)_{\msc{out}} &= \int_0^{\infty}\frac{\dd p}{2 \pi} \,
4331: \Big|d(p,\omega(p,M))\Big|^2 \, \frac{1} {2 \om(p,M)} \Big|
4332: \Psi_{\rm absorb} (\rho_0,t_0;p,\om(p,M))\Big|^2 \nonumber \\
4333: &\sim {1 \over M} \int_0^{\infty} {\dd p} \, e^{-\pi
4334: \left(1+\frac{1}{k}\right)p - \pi\sqrt{p^2+ 2k M^2}}~. \label{N M
4335: super out}
4336: \end{align}
4337: In the second lines, we have taken $M$ large, viz. $\omega \gg p \gg
4338: 1$, and keep the leading terms only. Thus, for each fixed but large
4339: $M$, the partial number distributions take the forms:
4340: \begin{align}
4341: \cN (M)_{\msc{in}} \sim \int_0^{\infty} \dd p \,
4342: \sigma_{\msc{in}}(p) e^{-\frac{1}{2}\beta_{\rm Hw} M} \qquad
4343: \mbox{and} \qquad N(M)_{\msc{out}} \sim \int_0^{\infty} \dd p \,
4344: \sigma_{\msc{out}}(p) e^{-\frac{1}{2}\beta_{\rm Hw} M}~, 
4345: \label{grey body}
4346: \end{align}
4347: where
4348: \begin{align}
4349:  \beta_{\rm Hw} = 2\pi \sqrt{2k}~,
4350: \label{Hawking temp super}
4351: \end{align}
4352: is the inverse Hawking temperature of the fermionic two-dimensional
4353: black hole. As discussed above, the radiation off the D-brane in
4354: NS5-brane background is analogous to the decay of excited D-brane in
4355: flat ambient space-time. 
4356: Indeed, asymptotic expression \eqref{grey
4357: body} suggests that open string 
4358: excitations of energy $M$ on 
4359: the rolling D0-brane are 
4360: populated as the distribution function $\exp
4361: (-{1 \over 2} \beta_{\rm Hw} M)$
4362: and decay into closed string
4363: radiation. 
4364: In this interpretation, the distribution function encodes
4365: change of available states for open string excitations on the
4366: D0-brane after emitting radiations of energy $M$. 
4367: Curiously,
4368: `effective temperature' 
4369: of the excited closed strings is set by the
4370: Hawking temperature of the nonextremal NS5-brane, not that of a
4371: black hole that would have been made 
4372: of the D0-brane. It is tempting
4373: to interpret this as indicating 
4374: that the D0-brane represents a class
4375: of possible excitation modes of the black NS5-brane. The closed
4376: string states of energy $M$ emitted 
4377: by the D0-brane are certainly coherent, 
4378: but according to this interpretation, they still can be
4379: recasted in effective thermal distribution set by the Hawking
4380: temperature of the two-dimensional black hole. 
4381: %%%%
4382: %%%%
4383: We will later discuss again the origin of such effective
4384: thermal behavior of the rolling D0-brane 
4385: from the viewpoints of Euclidean cylinder 
4386: amplitudes, extending the argument of \cite{Dijkgraaf:1992ba} 
4387: about the Hawking radiation in the purely closed string 
4388: background. 
4389: %%%%
4390: %%%%
4391: 
4392: 
4393: The functions
4394: $\sigma_{\rm in}$ and $\sigma_{\rm out}$ are 
4395: interpretable as the black hole `greybody' factors 
4396: for incoming and outgoing parts of the
4397: radiation. The factor 1/2 in the exponent of the Boltzmann
4398: distribution function reflects the fact that only left-right
4399: symmetric closed string states can appear 
4400: in the boundary states and the radiated closed string modes.
4401: 
4402: 
4403: 
4404: The `greybody factors' $\sigma_{*}(p)$ depend on the radial momentum
4405: $p$ exponentially, so the radiation distribution would be modified
4406: {\sl once} the radial momentum $p$ is integrated out. Below, we
4407: shall show this explicitly. We are primarily interested in keeping
4408: track of string world-sheet effects set by the value of the level
4409: $k$. We shall consider different ranges of the level $k$ separately,
4410: and focus on the asymptotic behaviors at large $M$ via the saddle
4411: point methods.
4412: 
4413: %%%%%%
4414: \begin{description}
4415:  \item[(i) \underline{$k > 1$}: ] \hfill\break
4416: 
4417: This is the case for the black NS5-brane background. Consider first
4418: the incoming part. Since $1- \frac{1}{k}> 0$, the dominant
4419: contribution in the $p$-integral arises from the saddle point:
4420: \begin{align}
4421: p \sim p_* = \frac{k-1}{\sqrt{1-\frac{1}{2k}}} M~.\nonumber
4422: \end{align}
4423: Substituting this to \eqref{N M super in}, we obtain
4424: \begin{align}
4425: \cN (M)_{\msc{in}} \sim e^{-2\pi M \sqrt{1-\frac{1}{2k}}} =
4426: e^{-\frac{1}{2}M \beta_{\rm Hg}}~, \label{eq:in}
4427: \end{align}
4428: up to pre-exponential powers of $M$. Taking account of the density
4429: of states $\rho(M) \sim e^{\frac{1}{2}M \beta_{\rm Hg}}$, we find
4430: that $\rho(M) \cN(M)_{\msc{in}}$ scales with powers of $M$, and is
4431: independent of $k$. More explicitly, for the black NS5-brane $\cM=
4432: SU(2)_{k} \times \mathbb{R}^5$, the incoming radiation
4433: distribution of the D$p$-brane parallel to the NS5-brane yields
4434: \begin{align}
4435: \cN(M)_{\msc{in}} &\sim {1 \over M} \int {\dd^{5-p} {\bf k}_{\perp}
4436: \over (2 \pi)^{5-p}} \int_0^\infty {\dd p} \, e^{\pi
4437: (1-\frac{1}{k})p-\pi\sqrt{p^2+2k(M^2+{\bf k}_{\perp}^2)}}
4438: \nonumber \\
4439: &\sim M^{2-\frac{{p}}{2}}\, e^{-2\pi M\sqrt{1-\frac{1}{2k}}} \ . \nonumber
4440: \end{align}
4441: Taking  account of the density of states $\rho(M) \sim
4442: M^{-3}e^{2\pi M\sqrt{1-\frac{1}{2k}}}$, the average radiation number
4443: distribution is given by
4444: \begin{align}
4445: \frac{\overline{\cN}_{\msc{in}}}{V_p} \sim \int^{M_{\rm D}} {\dd M
4446: \over M} \, M^{-\frac{p}{2}} \qquad \mbox{where} \qquad M_{\rm D} \sim {\cal
4447: O}({1 \over g_{\rm st}}) \ . \label{conL}
4448: \end{align}
4449: This result coincides with the computations of \cite{Lambert:2003zr,Karczmarek:2003xm}, and
4450: corroborates with the radion-tachyon correspondence. Interestingly,
4451: the incoming part of the radiation number distribution in the the
4452: nonextremal NS5-brane background is exactly the same as the
4453: distribution in the extremal NS5-brane background. Later, we shall
4454: examine carefully taking the extremal limit and its consequence in
4455: section \ref{sec:8-6}. As in the extremal case, \eqref{conL} implies that
4456: nearly all the D0-brane potential energy is released into closed
4457: string radiations before it falls into the black hole.
4458: 
4459: On the other hand, for the outgoing radiation, the far infrared $p
4460: \sim 0$ dominates the momentum integral. We thus obtain
4461: \begin{align}
4462: \cN (M)_{\msc{out}} \sim e^{-2\pi M \sqrt{\frac{k}{2}}} =
4463: e^{-\frac{1}{2}M \beta_{\rm Hw}}~, \nonumber
4464: \end{align}
4465: displaying effective thermal distribution set by the Hawking
4466: temperature. Taking account of the density of states,
4467: \begin{align}
4468: \rho(M) \cN (M)_{\msc{out}} \sim e^{\frac{1}{2}M \left(\beta_{\rm
4469: Hg}-\beta_{\rm Hw}\right)} = e^{2\pi M
4470: \left(\sqrt{1-\frac{1}{2k}}-\sqrt{\frac{k}{2}} \right)}~. \nonumber
4471: \end{align}
4472: This is ultraviolet finite for any $k$ since
4473: \begin{align}
4474: \left(1-\frac{1}{2k}\right) -\frac{k}{2} = -\frac{1}{2k}
4475: \left(k-1\right)^2 < 0~. \label{eq:ini}
4476: \end{align}
4477: 
4478: We thus conclude that the radiation number distribution is mostly in
4479: the incoming part:
4480: %
4481: \begin{align} \left. {{\cal N}_{\rm out}(M) \rho (M) \over {\cal N}_{\rm
4482: in}(M) \rho (M)} \right\vert_{\rm falling \,\, D0} \sim {e^{-{1
4483: \over 2} \beta_{\rm Hg} M } \over e^{-{1 \over 2} \beta_{\rm Hw} M}}
4484: = e^{2 \pi M \left( \sqrt{1 -{1 \over 2 k}} - \sqrt{k \over 2}
4485: \right)} \ll 1. \nonumber \end{align}
4486: %
4487: Intuitively, this may be understood as follows: for the absorbed
4488: boundary state, the boundary condition is such that the D0-brane
4489: flux is directed from past null infinity to the future horizon. This
4490: also corroborates the observation that $T_{t\rho}$-component of
4491: D0-brane's energy-momentum tensor is nonzero and increases
4492: monotonically as the D0-brane approaches the future horizon. The
4493: outgoing part of the distribution is exponentially small compared to
4494: the incoming part and exhibits effective thermal distribution at the
4495: Hawking temperature. Notice that, despite being so, this outgoing
4496: part has nothing to do with the Hawking radiation of the black hole.
4497: The latter is the feature of the background by itself. A priori, the
4498: outgoing radiation could be in a distribution characterized by a
4499: temperature different from the Hawking temperature. As mentioned
4500: above, it is tempting to interpret coincidence of the two
4501: temperatures as a consequence of maintaining equilibrium between the
4502: black NS5-brane and the D0-brane.
4503: 
4504: %%%
4505: \item[(ii) \underline{$\frac{1}{2} < k  \leq 1$}: ]
4506: \hfill\break
4507: 
4508: This is the regime which includes the conifold geometry at $k=1$.
4509: Since $1-\frac{1}{k} \leq 0$, the dominant contribution to the
4510: momentum integral is from $p \sim 0$, not only for the outgoing
4511: radiation but also for the incoming one. We thus obtain
4512: \begin{align}
4513: \cN(M)_{\msc{in}} \sim \cN (M)_{\msc{out}} \sim e^{-2\pi M
4514: \sqrt{\frac{k}{2}}} \equiv e^{-\frac{1}{2}M \beta_{\rm Hw}}~,
4515: \end{align}
4516: viz. both are in effective thermal distribution set by the Hawking
4517: temperature. All spectral moments are manifestly ultraviolet finite
4518: since, at large $M$, exponential growth of the density of the final
4519: closed string states is insufficient to overcome the suppression by
4520: the distribution. Thus,
4521: %
4522: \begin{align} {{\cal N}_{\rm out}(M) \rho(M) \over {\cal N}_{\rm in} (M) \rho
4523: (M)} \Big|_{\rm falling \, D0} \sim 1. \nonumber \end{align}
4524: %
4525: We interpret this as indicating that the D0-brane does not radiate
4526: off most of its energy before falling into the horizon.
4527: 
4528: %%%%
4529: \item[(iii) \underline{$k=\frac{1}{2}$} : ]
4530: \hfill\break
4531: 
4532: This special case corresponds to empty $\cM$. The two-dimensional
4533: background permits no transverse degrees of freedom of the string.
4534: The physical spectrum includes massless tachyon only, with $M=0$ and
4535: $\rho(M)=1$. We now have a crucial difference from the previous
4536: cases for the on-shell configurations. The radial momentum $p$ is
4537: fixed by the on-shell condition as $\om = \pm p$, so it should not
4538: be integrated over for the final states. Consequently, we cannot
4539: decompose the radiation distribution into incoming and outgoing
4540: radiations,
4541: %because of the non-vanishing interference term,
4542: and only the total distribution is physically relevant.
4543: %We also note that the canonically normalized energy is given as $E =
4544: %p = \om$ from the on-shell condition.
4545: 
4546: We thus obtain the following large $\om$ behavior of the radiation
4547: distribution:
4548: \begin{align}
4549: \cN(\om)\sim  e^{-2\pi \om} \equiv e^{-\om \beta_{\rm Hw}}~.
4550: \label{radiation 2D BH super}
4551: \end{align}
4552: Again, we have found effective thermal distribution at the Hawking
4553: temperature! Notice the absence of extra 1/2-factor in contrast to
4554: the previous regimes. This is not a contradiction. In the present
4555: case, the transverse oscillators are absent and the string behaves
4556: as a point particle. Again, the D0-brane does not radiate off most
4557: of its energy before falling across the black hole horizon.
4558: In the linear dilaton regime, the boundary states and possible connection with the matrix model for the two-dimensional type 0A/0B string theory have been discussed in \cite{Lapan:2005qz}. Given the phase transition we observed, however, the classical intuition of such ``rolling D-brane" in the two-dimensional noncritical string theory is rather questionable. It would be interesting to give an interpretation from the dual matrix models.
4559: \end{description}
4560: 
4561: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4562: 
4563: \subsubsection{radiation distribution in bosonic string theory}\label{sec:8-1-2}
4564: 
4565: The analysis for the bosonic string case proceeds quite the same
4566: route. The boundary state for the infalling D0-brane includes the
4567: string world-sheet correction factor
4568: $\Gamma\left(1+i\frac{p}{\kappa-2}\right)$, where again $\kappa$
4569: refers to the level of bosonic $SL(2;\br)/U(1)$ coset model. The
4570: on-shell condition now reads
4571: \begin{align}
4572: & -\frac{\om^2}{4\kappa} + \frac{p^2}{4(\kappa-2)}+
4573: \frac{1}{4(\kappa-2)} + \Delta_\cM = 1~, \label{on-shell bosonic}
4574: \end{align}
4575: where $\Delta_\cM$ denotes the conformal weight in the $\cM$-sector.
4576: This is solved by
4577: \begin{align}
4578: & \om \equiv \om(p,M) = \sqrt{\frac{\kappa}{\kappa-2}p^2+2\kappa
4579: M^2} \qquad \mbox{where} \qquad M^2 \equiv 2 \left(\Delta_\cM +
4580: \frac{1}{4(\kappa-2)} - 1\right)~.
4581: \end{align}
4582: The partial radiation number distribution at large $M$ limit is
4583: given by:
4584: \begin{align}
4585: & \cN(M)_{\msc{in}}
4586: %= \int_0^{\infty}\frac{dp}{\om(p,M)}\,
4587: %\left| \Psi(p,\om(p,M))\right|^2 \nn
4588:  \sim  {1 \over M} \int_0^{\infty} {\dd p} \,
4589: e^{+\pi \left(1- \frac{1}{\kappa-2}\right)p - \pi
4590: \sqrt{\frac{\kappa}{\kappa-2} p^2 + 2\kappa M^2}}~.
4591: \label{N M bosonic in} \\
4592: & \cN(M)_{\msc{out}}
4593: %= \int_0^{\infty}\frac{dp}{\om(p,M)}\,
4594: %\frac{\cosh^2 \pi\left(\om (p,M)- \frac{p}{\cQ}\right)}
4595: %{\cosh^2 \pi\left(\om (p,M)+ \frac{p}{\cQ}\right)}
4596: %\left| \Psi(p,\om(p,M))\right|^2 \nn
4597:  \sim  {1 \over M} \int_0^{\infty} {\dd p} \,
4598: e^{-\pi \left(1+ \frac{1}{\kappa-2}\right)p
4599: - \pi \sqrt{\frac{\kappa}{\kappa-2}p^2+ 2\kappa M^2}}~.
4600: \label{N M bosonic out}
4601: \end{align}
4602: Thus, as in the superstring case, there can arise several distinct
4603: behaviors depending on how stringy the background is.
4604: 
4605: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4606: 
4607: \begin{description}
4608:  \item[(i) \underline{$\kappa > 3$}: ]
4609: \hfill\break
4610: %
4611: Consider first the incoming radiation part. Since
4612: $1-\frac{1}{\kappa-2}
4613: > 0$, the dominant contribution to the momentum integral in \eqref{N M
4614: bosonic in} is from the saddle point
4615: \begin{align}
4616: p \sim p_* = \frac{\kappa-3} {\sqrt{2-\frac{1}{2(\kappa-2)}}} M~.
4617: \nonumber
4618: \end{align}
4619: We thus obtain, up to pre-exponential powers of $M$,
4620: \begin{align}
4621: \cN(M)_{\msc{in}} \sim e^{-2\pi M \sqrt{2-\frac{1}{2(\kappa-2)} } }
4622: \, = \, e^{-\frac{1}{2} M \beta_{\rm Hg}}~, \nonumber
4623: \end{align}
4624: where $\beta_{\rm Hg}$ denotes the Hagedorn temperature of the
4625: bosonic string theory \eqref{Hagedorn bosonic}.
4626: %The density of states for the bosonic model is given as
4627: %$\rho(M) \sim e^{\frac{1}{2}\beta_{Hg}M} =
4628: %e^{2\pi \sqrt{2\left(1-\frac{\cQ^2}{8}\right)}M}$,
4629: In this way, we again find the power-law behavior of $\rho(M) \cN
4630: (M)_{\msc{in}}$ at large $M$, independent of the level $\kappa$.
4631: 
4632: For the outgoing radiation part, again the $p\sim 0$ dominates the
4633: momentum integral in \eqref{N M bosonic out}. The result is
4634: \begin{align}
4635: \cN (M)_{\msc{out}} \sim e^{-2 \pi M \sqrt{\frac{\kappa}{2}}} =
4636: e^{-\frac{1}{2} M \beta_{\rm Hw}}~. \nonumber
4637: \end{align}
4638: Here,
4639: \begin{align}
4640:  \beta_{\rm Hw} \equiv 2\pi \sqrt{2\kappa} \nonumber
4641: \end{align}
4642: is the Hawking temperature of the bosonic two-dimensional black
4643: hole. We then obtain
4644: \begin{align}
4645: \rho(M) \cN (M)_{\msc{out}} \sim e^{\frac{1}{2}\left(\beta_{\rm
4646: Hg}-\beta_{\rm Hw}\right)M} = e^{2\pi M
4647: \left[\sqrt{2-\frac{1}{2(\kappa-2)}} -\sqrt{\frac{\kappa}{2}}
4648: \right] }~. \nonumber
4649: \end{align}
4650: As in the superstring case, the exponent is always negative
4651: definite:
4652: \begin{align}
4653: \left(2-\frac{1}{2(\kappa-2)}\right)
4654: -\frac{\kappa}{2} =
4655: -\frac{\left(\kappa-3\right)^2}{2(\kappa-2)}
4656:  \leq  0~. \nonumber
4657: \end{align}
4658: so the outgoing radiation distribution (as well as spectral moments)
4659: is manifestly ultraviolet finite.
4660: 
4661: Physical interpretation of the above results is the same as the
4662: superstring case: The D0-brane falling into the black hole has
4663: nonzero component $T_{t\rho}$ of the energy-momentum tensor, and
4664: entails that dominant part of the closed string radiation is
4665: incoming toward the future horizon. The outgoing part of the
4666: radiation is exponentially suppressed, and is in effective thermal
4667: distribution set by the Hawking temperature. Again, this
4668: distribution is distinct from the Hawking radiation of the
4669: two-dimensional black hole. As for the fermionic string, the
4670: branching ratio is exponentially suppressed.
4671: 
4672: %%%
4673: \item[(ii) \underline{$\frac{9}{4} < \kappa \leq 3$}: ]
4674: \hfill\break
4675: %
4676: In this regime, $1- \frac{1}{\kappa-2} < 0$ and the momentum
4677: integrals for both incoming and outgoing radiation distributions are
4678: dominated by $p \sim 0$:
4679: \begin{align}
4680: \cN(M)_{\msc{in}} \sim \cN(M)_{\msc{out}} \sim  e^{-2 \pi M
4681: \sqrt{\frac{\kappa}{2}}} \equiv e^{-\frac{1}{2}M \beta_{\rm Hw}}~.
4682: \nonumber
4683: \end{align}
4684: Both are in effective thermal distribution at the Hawking
4685: temperature, and all spectral moments are manifestly ultraviolet
4686: finite since, at large $M$, the growth of the density of state does
4687: not overcome the suppression by the distribution. The branching
4688: ratio remains order unity.
4689: 
4690: %%%%
4691: \item[(iii) \underline{$\kappa=\frac{9}{4}$} : ]
4692: \hfill\break
4693: %
4694: This is the most familiar situation: black hole in two-dimensional
4695: bosonic string theory, originally studied in \cite{Witten:1991yr,Elitzur:1991cb,Mandal:1991tz,Bars:1990rb}. The
4696: physical spectrum of closed string consists only of the massless
4697: tachyon, so we again need to set $M=0$ and $\rho(M)=1$. The
4698: calculation is slightly more complicated than the supersymmetric
4699: case: The canonically normalized energy is
4700: \begin{align}
4701:  E = \frac{\sqrt{2}}{3}\om = \sqrt{2}p~, \nonumber
4702: \end{align}
4703: so we obtain
4704: \begin{align}
4705: & \cN (E) \sim  e^{\pi \left(1-\frac{1}{1/4}\right) p -\pi
4706: \sqrt{\frac{9/4}{1/4}}p} = e^{-3 \sqrt{2}\pi E} \equiv e^{- E
4707: \beta_{\rm Hw}}~. \nonumber
4708: \end{align}
4709: It again shows effective thermal distribution of the radiated closed
4710: string modes at the Hawking temperature: $\beta_{\rm Hw} = 2\pi
4711: \sqrt{2\kappa} = 3\pi \sqrt{2}$.
4712: 
4713: \end{description}
4714: 
4715: 
4716: 
4717: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4718: 
4719: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4720: \subsubsection{radiation distribution for emitted or time-symmetric boundary states}\label{sec:8-1-3}
4721: 
4722: The closed string radiations for the other types boundary states,
4723: viz. the `emitted' \eqref{emitted D0} or the `symmetric'
4724: \eqref{symmetric D0} D0-branes, can be studied analogously.
4725: 
4726: For the emitted D0-brane boundary state \eqref{emitted D0}, by the
4727: time-reversal, we should observe the radiation distribution at the
4728: far past: $t\sim -\infty$. The relevant decomposition corresponding
4729: to \eqref{as U} is given by (assuming $\om>0$, $p>0$)
4730: \begin{align}
4731:  && V^p_{\om}(\rho,t) \sim e^{i\om \ln \rho-i\om t}
4732:   + d^*(p,\om) e^{-\rho} e^{-ip\rho -i\om t} ~,
4733: \label{as V}
4734: \end{align}
4735: where the first term is supported near the past horizon
4736: %($\rho \sim 0$)
4737: and the second term corresponds to the incoming wave from the null
4738: infinity.
4739: %($\rho \sim +\infty$)
4740: Obviously we find precisely the same behavior of the radiation
4741: distribution as the absorbed D0-brane once the role of `in' and
4742: `out' states are reversed. So, for $k>1$, ${\cal N}(M)_{\rm in} \sim
4743: \exp ( - {1 \over 2} \beta_{\rm Hw} M)$ while ${\cal N}(M)_{\rm out}
4744: \sim \exp (-{1 \over 2} \beta_{\rm Hg} M)$ and, for $ 1 \ge k >
4745: 1/2$, ${\cal N}(M)_{\rm in}$, ${\cal N}(M)_{\rm out} \sim \exp (-{1
4746: \over 2} \beta_{\rm Hw} M)$.
4747: 
4748: Consider next the boundary state describing D0-brane in symmetric
4749: boundary condition \eqref{symmetric D0}. Recalling the relations
4750: \eqref{rel disc amp}, one finds that the radiation rates are simply
4751: obtained by adding contributions from `absorbed' and `emitted'
4752: D0-brane boundary states. Thus, the radiation distributions behave as
4753: ${\cal N}(M)_{\rm in}$, ${\cal N}(M)_{\rm out} \sim \exp (-{1 \over
4754: 2} \beta_{\rm Hg} M)$ for $k>1$ and the dependence on Hawking
4755: temperature disappeared.\footnote{Dependence on the Hawking
4756: temperature exponentially suppressed, so completely negligible
4757: compared to other power-suppressed subleading terms.} We then find
4758: that the `detailed balance' $\cN(M)_{\msc{in}} = \cN(M)_{\msc{out}}$
4759: is obeyed. This is as expected since the boundary state
4760: \eqref{symmetric D0} is defined so that it keeps the time-reversal
4761: symmetry and the one-particle state unitarity manifest.
4762: 
4763: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4764: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4765: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4766: \subsubsection{revisit to the radiation distribution from 
4767: thermal sting propagator}\label{sec:8-1-4}
4768: 
4769: 
4770: To close this section we discuss the radiation 
4771: distribution from a different angle. 
4772: Although the argument given here would be somewhat 
4773: heuristic, it is quite helpful to grasp physical intuition 
4774: and to understand where the thermal-like behavior 
4775: of closed string radiation comes from. This argument 
4776: is much like the one given in \cite{Dijkgraaf:1992ba}, where 
4777: the Hawking radiation of 2-dimensional black-hole is discussed
4778: by the closed string thermal propagator. In a sense, 
4779: our discussion is an extension of it to the open string sector.
4780: 
4781: 
4782: We start with the (thermal) cylinder amplitude 
4783: for the D1-brane on the Euclidean cigar \eqref{clss2'}
4784:    \footnote{To be more precise, we consider
4785:    the fermionic black-hole of level $k$ 
4786:     and focus on the space-time bosons. 
4787:     If considering the space-time fermions, the thermal KK 
4788:     momentum should be half integer $n\in 1/2 + \bz$, rather
4789:     than $n\in \bz$, which leads to the fermionic distribution 
4790:    $1/(e^{\beta_{\msc{Hw}}\om_{p,M}}+1) $ instead of 
4791:     $1/(e^{\beta_{\msc{Hw}}\om_{p,M}}-1)$ in the following 
4792:     argument.}. 
4793: This is approximately evaluated as (we omit the parameters
4794: $\rho_0$, $\theta_0$ for simplicity)
4795: \begin{align}
4796: & \cA_{\msc{cylinder}}^{(E)} \cr &= \int_0^{\infty} dT\, 
4797:  {}_{D1} \bra{B} e^{-\pi T H^{(c)}} \ket{B}_{D1} 
4798:  \approx  \sum_M \sum_{n\in \bsz} \int dp \, \frac{1}
4799: {p^2+ \left(\frac{2\pi n}{\beta_{\msc{Hw}}}
4800: \right)^2+M^2} \, \sqrt{\rho(M)}
4801:   \left|\Psi_{D1}(p,n)\right|^2~ \cr
4802: & = \frac{\beta_{\msc{Hw}}}{2\pi} \sum_M  
4803: \int dpdq \, \sqrt{\rho(M)}
4804: \frac{\left|\Psi_{D1}\left(p,\frac{\beta_{\msc{Hw}}q}{2\pi}
4805: \right)\right|^2}{p^2+q^2+M^2}\,
4806: \left(1+ \sum_{m\in \bsz_{>0}}e^{i \beta_{\msc{Hw}} m q}
4807:  + \sum_{m\in \bsz_{>0}}e^{-i \beta_{\msc{Hw}} m q}
4808: \right)~.
4809: \label{thermal cylinder}
4810: \end{align}
4811: Here $p$ is the radial momentum and $n$ is the KK
4812: momentum along the asymptotic circle of cigar 
4813: (thermal circle). 
4814: %The absence of winding is obvious because of the boundary 
4815: %condition for the D1-brane on the cigar 
4816: %(Dirichlet in the asymptotic region).
4817: $M$ is again the transverse mass in the $\cM$-sector and 
4818: $\rho(M)$ is the density of closed string states.
4819: $\beta_{\msc{Hw}}\equiv 2\pi \sqrt{2k}$  again denotes the 
4820: inverse Hawking temperature. 
4821: Let us try to Wick rotate it by 
4822: the contour deformation of $q$-integration 
4823: in the similar manner to \cite{Dijkgraaf:1992ba}. 
4824: Setting $q=i \om$\footnote
4825:     {Here,  $\omega$,  $p$ are  normalized 
4826:     as $L_0 = - \frac{1}{2} \om^2 + \frac{1}{2}p^2
4827:    + \cdots$, rather than
4828:     $L_0 = -\frac{1}{4k} \om^2 + \frac{1}{4k} p^2 + \cdots$.},
4829: $\cA^{(L)}_{\msc{cylinder}} = -i \cA^{(E)}_{\msc{cylinder}}$,
4830: we formally obtain 
4831: \begin{align}
4832:  & \cA^{(L)}_{\msc{cylinder}} \cr & \approx
4833: \frac{\beta_{\msc{Hw}}}{2\pi} \sum_M \int dp\, 
4834: \sqrt{\rho(M)}\, \left\lb 
4835: \int d\om \, \frac{
4836: \left|\Psi_{D1}\left(p,\frac{i \beta_{\msc{Hw}}\om}{2\pi}
4837: \right)
4838: \right|^2
4839: }{p^2+M^2-\om^2+i\ep}
4840: -\frac{2\pi i }{\om_{p,M}} 
4841: \frac{\left|\Psi_{D1}\left(p,
4842: \frac{i \beta_{\msc{Hw}}\om_{p,M}}
4843: {2\pi}\right)\right|^2}
4844: {e^{\beta_{\msc{Hw}} \om_{p,M}}-1}
4845: \right\rb ~, 
4846:  \label{evaluation AL}
4847: \end{align}
4848: where $\om_{p,M} \equiv \sqrt{p^2+M^2}$ 
4849: is the on-shell energy  and we here used 
4850: $$
4851: \left|\Psi_{D1}\left(p,-\frac{i \beta_{\msc{Hw}}\om_{p,M} }
4852: {2\pi}\right)\right|^2 = 
4853: \left|\Psi_{D1}\left(p,\frac{i \beta_{\msc{Hw}}\om_{p,M} }
4854: {2\pi}\right)\right|^2~. 
4855: $$
4856: Because we have 
4857: $ \left|\Psi_{D1}\left(p,\frac{i \beta_{\msc{Hw}}\om}{2\pi}
4858: \right)
4859: \right|^2  \propto 
4860: e^{\frac{1}{2}\beta_{\msc{Hw}} |\om|}$, 
4861: the first term (including the Feynman propagator) 
4862: gives a UV divergent contribution. 
4863: This is not surprising and shows the reason why the naive 
4864: Wick-rotation of \eqref{clss2'} does not work. 
4865: The second term shows a `thermal-like' form and 
4866: actually contributes to the imaginary part 
4867: of cylinder amplitude we are interested in. 
4868: It gives the expected behavior;
4869: \begin{align}
4870: \Im \, \cA_{\msc{thermal}}^{(L)} &\propto 
4871:   \frac{1}{\om_{p,M}} 
4872: \frac{1}{e^{\beta_{Hw}\om_{p,M}}-1} \, \sqrt{\rho(M)} 
4873:  \left|\Psi_{D1}\left(p,\frac{i \beta_{\msc{Hw}}\om}{2\pi}
4874:  \om\right)\right|^2 \cr
4875: &\sim \frac{\sqrt{\rho(M)} \sigma(p)}{\om_{p,M}}\,
4876:  e^{-\frac{1}{2}\beta_{\msc{Hw}} \om_{p,M}}~,
4877: \label{thermal behavior}
4878: \end{align}
4879: %because we have 
4880: %$\left|\Psi_{D1}(p,iR\om)\right|^2 \sim \sigma(p) 
4881: %e^{\frac{1}{2}\beta_{Hw} \om}$. 
4882: which reproduces the previous results \eqref{grey body},
4883: including the correct grey body factor $\sigma(p)$.
4884: %based on the exact boundary states 
4885: %for the Lorentzian D0-branes. 
4886: Recall that, in our construction of Lorentzian boundary 
4887: states, the presence of the damping factor was crucial, 
4888: which reads as 
4889: $\frac{\sinh \pi \sqrt{2k} p}
4890: {\cosh \left\lb \pi \sqrt{\frac{k}{2}}(p+\om) \right\rb
4891: \cosh \left\lb \pi \sqrt{\frac{k}{2}}(p-\om)\right\rb} $ 
4892: in the convention here. 
4893: This factor shows the same asymptotic behavior 
4894: in the large $\om$ or large $p$ region as 
4895: the Boltzmann distribution functions
4896: $1/(e^{\beta_{\msc{Hw}}\om}\pm 1)$.
4897: In this sense, our Wick-rotation of boundary states 
4898: would be roughly identified with the procedure 
4899: which keeps only the second term in \eqref{evaluation AL}, 
4900: suggesting the origin of the thermal-like distribution
4901: derived from our Lorentzian boundary states. 
4902: 
4903: 
4904: Another helpful argument is achieved by starting with 
4905: the open string channel of thermal cylinder amplitude 
4906: \eqref{thermal cylinder}. 
4907: Let us focus on the asymptotic region $\rho \gg 0$
4908: for simplicity, 
4909: in which the hairpin D1-brane \eqref{clss2'} appears just
4910: as two halves of $D1$-$\bar{D}1$ system, which are Dirichlet
4911: along the thermal circle, 
4912: (so, identified as the `$sD$-$s\bar{D}$ system' 
4913: \cite{Strominger:2002pc}) as pointed out in \cite{Nakayama:2004yx}. 
4914: In this set up, by a simple kinematical reason, 
4915: we find {\em on-shell} closed string states in the cylinder
4916: amplitude, while only {\em off-shell} states in the open 
4917: string channel.  As discussed {\em e.g.} 
4918: in \cite{Sugawara:2002rs,Sugawara:2003xt}, using the 
4919: modular transformation, we can show the thermal distribution of 
4920: {\em physical} closed string states emitted/absorbed by 
4921: the $sD$-$s\bar{D}$ system is captured by   
4922: the {\em unphysical} open string winding modes along 
4923: the thermal circle\footnote
4924:    {This is a simple extension of the standard argument 
4925:     of the thermal toroidal partition functions 
4926:     \cite{Polchinski:1985zf,Sathiapalan:1986db,Kogan:1987jd,O'Brien:1987pn,Atick:1988si}. For instance, the Hagedorn behavior
4927:     is interpretable as the tachyonic instability due to the 
4928:      {\em unphysical} winding modes along the thermal circle.
4929:      }. 
4930: Especially, the unit of winding energy should 
4931: determine the temperature of thermal distribution 
4932: of closed string states coupled with 
4933: the $sD$-$s\bar{D}$ system. 
4934: In the present case it is identified with the interval 
4935: of the hairpin 
4936: $(= \frac{1}{2}\beta_{\msc{Hw}})$, which is just associated 
4937: to the $D1$-$\bar{D}1$ open string. 
4938: (Note that, taking suitably the GSO projection 
4939: into account, we can find the zero winding
4940: modes, {\em i.e.} the $D1$-$D1$ or $\bar{D}1$-$\bar{D}1$ 
4941: strings, are canceled out. See \cite{Nakayama:2004yx}.)
4942: This is the simplest explanation of why we get 
4943: the thermal-like distribution 
4944: $\propto e^{-\frac{1}{2}\beta_{\msc{Hw}}\om_{p,M}}$
4945: from the cylinder amplitude \eqref{thermal cylinder}.
4946: 
4947: 
4948: Curiously, all the {\em regular} solutions 
4949: of $D0$-brane motion are just straight lines 
4950: in the Kruscal coordinates, and thus Wick-rotated 
4951: to the hairpin profiles with the {\em same} 
4952: interval; $\frac{1}{2}\beta_{\msc{Hw}}$. This fact
4953: leads us to the same thermal-like behaviors \eqref{grey body} 
4954: characterized by the Hawking temperature (before integrating
4955: $p$ out)\footnote
4956:    {One might ask why the D0-brane motion with different 
4957:     `temperature' is not considered. However,  
4958:     such D0-branes correspond to singular 
4959:     hairpin profiles and thus 
4960:     do to singular Lorentzian trajectories. 
4961:     These cannot be the solutions of DBI action of D0-branes
4962:     by the divergence of the velocity at the singular points. 
4963:     Quite interestingly, this feature is similar to 
4964:     the original Hawking's idea : requiring the smooth 
4965:     Euclidean geometry, we can fix the particular asymptotic 
4966:     periodicity of Euclidean time, which yields 
4967:     the temperature characterizing the radiation 
4968:     from black-hole.}, 
4969: as is already pointed out.  
4970: 
4971: 
4972: 
4973: 
4974: 
4975: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4976: 
4977: \subsection{Radiation out of rolling D-brane from open string viewpoint}\label{sec:8-2}
4978: \subsubsection{open string channel viewpoint}\label{sec:8-2-1}
4979: What is the nature of the ultraviolet behavior of the emission
4980: number $\overline{\cal N}$ and how is it compared to the decay of
4981: rolling D-brane? To answer these, we shall now recast \eqref{ImZ 0} in
4982: the open string channel, following technical procedures considered
4983: in section \ref{sec:5-2} and appendix of \cite{Karczmarek:2003xm}.
4984: 
4985: In the closed string channel, the closed string radiation has been computed as  (see section \ref{sec:8-1})
4986: \begin{align} \overline{\cal N} = N^2_{\rm NS} \sum_M \sqrt{\rho^{(c)}(M)}
4987: \int_0^{\infty} {\rmd p \over 2 \pi} \, \frac{1}{2\om} {\cal P}(p,
4988: \omega) ~, \label{ImZ 0}\end{align}
4989: %
4990: where $N_{\rm NS}$ is an appropriate numerical factor, $\om =
4991: \sqrt{p^2 + 2kM^2}$ is the on-shell energy of the emitted closed
4992: string, and
4993: %
4994: \begin{align} {\cal P}(p, \omega) \equiv \left|\Psi(p,\om)\right|^2
4995:  = \frac
4996: {\sinh\pi\sqrt{\frac{\alpha'}{2}} p} {\left(\cosh \pi
4997: \sqrt{\frac{\alpha'}{2}} p + \cosh \pi \sqrt{\frac{\alpha'}{2}} \om \right)
4998: \sinh \frac{\pi}{k}\sqrt{\frac{\alpha'}{2}}  p }~ \label{powerspec}
4999: \end{align}
5000: is the transition probability.
5001: 
5002: We begin with expanding the transition probability ${\cal P}(p,\om)$ 
5003: of the D0-brane \eqref{powerspec} in power series of contribution of
5004: imaginary branes:
5005: %
5006: \begin{align}
5007:  & {\cal P}(p,\om) = \sum_{n=1}^{\infty}
5008: a_n(p) e^{-\pi n\omega\sqrt{\frac{\alpha'}{2}}} ~,
5009: \label{expansion}
5010: \\
5011:  & a_n(p)= 2(-1)^{n+1}
5012: \frac{\sinh\left(\pi n \sqrt{\frac{\alpha'}{2}} p \right)} {\sinh
5013: (\frac{\pi}{k}\sqrt{\frac{\alpha'}{2}} p)}~. \label{a n}
5014: \end{align}
5015: %
5016: %Notice that the power series
5017: %\eqref{expansion} is convergent around
5018: %the on-shell point $\om = \om_{p,M}$.
5019: %Using the elementary formula
5020: %
5021: %$\dsp \int_{-\infty}^{\infty} {\rmd k_0 \over 2 \pi} \,
5022: %\frac{e^{\frac{2\pi i n}{Q}\sqrt{\frac{\alpha'}{2}} k_0}}{k_0^2+p^2+M^2} =
5023: %\frac{1}{2 \om}e^{-\frac{2\pi n}{Q}
5024: %\sqrt{\frac{\alpha'}{2}} \om} $,
5025: As before,
5026: we parametrically rewrite \eqref{ImZ 0} as
5027: %
5028: \begin{align}
5029: \overline{\cal N} &=
5030: %\sum_M \sqrt{\rho^{(c)}(M)}
5031: %&\times& \int_0^{\infty}{\rmd p \over 2 \pi}
5032: % \sum_{n=1}^{\infty} \int_{-\infty}^{\infty} {\rmd k_0 \over 2 \pi} \,
5033: %  a_n(p) \frac{e^{2\pi i n \frac{k_0}{Q}}}{k_0^2+p^2+M^2}
5034: %\nn
5035:  N_{\rm NS}^2 \sum_M \sqrt{\rho^{(c)}(M)} \nn
5036:  %
5037:  &\times \int_0^{\infty} {\rmd p
5038: \over 2
5039:  \pi\sqrt{2k}} \sum_{n=1}^{\infty} \int_{-\infty}^{\infty}
5040: {\rmd k_0 \over 2 \pi \sqrt{2k}}\,
5041:  \int_0^{\infty} \frac{\alpha'}{2} \rmd t_c \,
5042:   a_n(p) e^{\pi i n \sqrt{\frac{\alpha'}{2}} k_0}
5043:  e^{-2\pi t_c \frac{1}{4}\alpha'\left(\frac{k_0^2+p^2}{2k}+M^2\right)}~, \qquad
5044: \label{ImZ 1}
5045: \end{align}
5046: %
5047: by introducing the Schwinger parameter $t_c$
5048: in the closed string channel.\footnote
5049:   {Strictly speaking, we could have the closed string tachyon
5050:    $M^2<0$, and the rewriting \eqref{ImZ 1} would not be
5051:    completely correct due to the infrared divergence. We can
5052:    avoid this difficulty by considering the GSO projected
5053:    amplitude. We are concerned with the large $M$ asymptotics,
5054:    so shall go on ignoring it to avoid unessential complexity.
5055:    }
5056: 
5057: 
5058: We now evaluate each contribution separately. Begin with the sum
5059: over the transverse mass $M$. By definition, the sum gives modular
5060: invariant cylinder amplitude of the $\cM$-sector:
5061: \begin{align}
5062: \sum_M \sqrt{\rho^{(c)}(M)} e^{-2\pi t_c \frac{\alpha'}{4} M^2} &=
5063: Z^{(c)}_{\cM}(q_c) \qquad \mbox{where} \qquad q_c = e^{- 2 \pi t_c}
5064: \nn &= Z^{(o)}_{\cM}(q_o) \qquad \mbox{where} \qquad q_o = e^{- 2
5065: \pi t_o} \quad (t_o \equiv 1/t_c)
5066: \end{align}
5067: by applying the standard open-closed duality and expressing the
5068: result in terms of the open string Schwinger parameter $t_o$.
5069: 
5070: 
5071: The amplitude $Z^{(o)}_{\cM}(t_o)$ asymptotes at large $t$ to
5072: (corresponding to the ultraviolet behavior in the closed string
5073: channel):
5074: \begin{align}
5075:  Z_{\cM}^{(o)}(t_o) \sim t_o^{\gamma}\,
5076: e^{2\pi t_o \cdot \frac{c_{\cM}}{24}}
5077:  = t_o^{\gamma} \, e^{\pi t_o \left(1-\frac{1}{2k}\right)} \qquad
5078:  \mbox{for} \qquad t_o \, \rightarrow\, +\infty.
5079: \end{align}
5080: Here, the exponent $\gamma$ is determined by the number of
5081: non-compact Neumann directions
5082: in the $\cM$-sector. Such details,
5083: however, are not relevant for our discussions.
5084: 
5085: The Gaussian integral over $k_0$ is readily evaluated,
5086: resulting in
5087: \begin{align}
5088:  & \overline{\cal N}
5089: = N^2_{\rm NS} \sqrt{\frac{\alpha'}{2}} \int_0^{\infty} {\rmd p \over 2 \pi\sqrt{2k}}
5090: \sum_{n=1}^{\infty}
5091:   \int_0^{\infty} \frac{\rmd t_o}{t_o^2} \sqrt{t_o} \,
5092:    a_n(p) \, e^{-\pi t_o \frac{n^2k}{2} -\frac{2\pi}{t_o} \frac{\alpha'}{4}\frac{p^2}{2k}}
5093:   \cdot Z^{(o)}_{\cM}(t_o)~.
5094: \end{align}
5095: %%%
5096: The $k_0$-integral yields the Boltzmann factor with the temperature
5097: determined by the Euclidean periodicity
5098: ($1/Q$ in our case) for the `hairpin brane'
5099: \cite{Ribault:2003ss,Eguchi:2003ik,Ahn:2003tt,Lukyanov:2003nj}, which is the Euclidean rotation of the rolling
5100: D-brane, as clarified in \cite{Nakayama:2005pk,Kutasov:2005rr}.
5101: This is essentially the same as the standard argument
5102: for thermal tachyon in the thermal string theory
5103: \cite{Polchinski:1985zf,Sathiapalan:1986db,Kogan:1987jd,O'Brien:1987pn,Atick:1988si}.
5104: %%%
5105: 
5106: 
5107: Our goal is to re-express the rate \eqref{ImZ 0} in the open string
5108: channel, so we shall Fourier transform the closed string momentum
5109: $p$ to the open string momentum $p'$. This requires a careful
5110: treatment, because the momentum-dependent coefficients $a_n(p)$ in
5111: \eqref{a n} could be exponentially growing functions. In such cases,
5112: the Fourier transform may not exists in a naive sense. We start with
5113: the identity:
5114: \begin{align}
5115:  & e^{-2\pi t_c \cdot \frac{1}{4} \alpha' \frac{p^2}{2k}} = \sqrt{t_o} \int_{\mathbb{R}+i\xi}
5116:  \sqrt{\frac{\alpha'}{2}}\frac{\rmd p'}{\sqrt{2k}}\,
5117: e^{-2\pi t_o \cdot \frac{1}{4} \alpha' \frac{p^{'2}}{2k}+2\pi i \cdot
5118: \frac{1}{2} \alpha' \frac{p p'}{2k}} \qquad \mbox{for} \qquad \xi \in \br~.
5119: \label{gauss xi}
5120: \end{align}
5121: In the $p$-integral, the function $e^{2\pi i \frac{1}{2} \alpha' \frac{p
5122: p'}{2k}}$ works as a damping factor and renders the integral finite if
5123: the parameter $\xi$ is chosen suitably. For later convenience, we
5124: shall decompose $a_n(p)$ as
5125: \begin{align}
5126:  & a_n(p) = a_n^+(p)-a_n^-(p)~, \nn
5127:  & a^{\pm}_n(p) \equiv (-1)^{n+1}
5128: \frac{e^{\pm\pi n \sqrt{\frac{\alpha'}{2}} p}}
5129:    {\sinh (\pi \frac{\pi}{k} \sqrt{\frac{\alpha'}{2}})}~.
5130: \label{a pm}
5131: \end{align}
5132: %
5133: %We first evaluate the integral
5134: %$\dsp \int_{-\infty}^{\infty}dp\,
5135: %a^+_n(p) e^{-\pi T p^2}$. Using \eqref{gauss xi},
5136: %we obtain
5137: %\footnote
5138: %  {Here, we are temporarily shifting
5139: %  the contour as $\br \, \rightarrow\, \br-i0$
5140: %  to avoid the pole $p=0$.
5141: %   We eventually restore it back to $\br$ {\sl after}
5142: %  taking the difference
5143: %  $a_n(p) \equiv a_n^+(p)-a_n^-(p)$.
5144: %  Also, if we choose the different contour shift
5145: %  $\br+i0$, \eqref{p-integral} should be
5146: %  replaced with
5147: %$$
5148: %\int_{\bsr + i0} \rmd p \frac{e^{\frac{2\pi n}{Q}p}}
5149: % {\sinh (\pi Q p)} e^{2\pi i p p'}
5150: % = {-i \over Q} \frac
5151: %{ e^{-\pi \left(\frac{p'}{Q}-i \frac{n}{Q^2}\right)}} {\cosh \pi
5152: %\left(\frac{p'}{Q}-i\frac{n}{Q^2}\right)}~.
5153: %$$
5154: %  However, the final result \eqref{ImZ final} remains intact. }
5155: %\begin{align}
5156: % && \int_{\bsr-i0} \rmd p\, a^+_n(p) e^{-\pi T p^2}
5157: %= \sqrt{t} \int_{\bsr-i0}\rmd p \int_{\bsr+i\xi_n^+} \rmd p'\,
5158: %  a^+_n(p) e^{2\pi i p p' - \pi t p^{'2}}~.
5159: %\label{evaluation 1}
5160: %\end{align}
5161: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5162: %The functions $a^+_n(p)$ behaves asymptotically as
5163: %\begin{align}
5164: % && a^+_n(p) \sim
5165: %\left\{
5166: %\begin{array}{ll}
5167: % 2(-1)^{n+1} e^{2\pi \left(\frac{n}{Q}-\frac{Q}{2}\right)p}&
5168: % ~~ p\,\rightarrow\, +\infty \\
5169: % 2(-1)^{n+1} e^{2\pi \left(\frac{n}{Q}+\frac{Q}{2}\right)p}&
5170: %  ~~ p\,\rightarrow\, -\infty
5171: %\end{array}
5172: %\right.
5173: %\end{align}
5174: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5175: Observing the asymptotic behavior of the coefficients $a^+_n(p)$, we
5176: readily find that the closed string channel momentum integral $\dsp
5177: \int \rmd p\, a^+_n(p) \, e^{2\pi i \cdot \frac{1}{2} \alpha' \frac{p p'}{2k}}$
5178: is well-defined as long as $\xi^+_n$ is chosen within the range
5179: $(nk-1)<\sqrt{\frac{\alpha'}{2}}\xi^+_n <
5180: (nk+1)$. We can then safely exchange the order
5181: of the integrals.
5182: %The $p$-integral is easily evaluated as
5183: %\begin{align}
5184: % && \int_{\bsr - i0} \rmd p \frac{e^{\frac{2\pi n}{Q}p}}
5185: % {\sinh (\pi Q p)} e^{2\pi i p p'}
5186: % = {i \over Q} \frac
5187: %{e^{\pi \left(\frac{p'}{Q}-i \frac{n}{Q^2}\right)}} {\cosh \pi
5188: %\left(\frac{p'}{Q}-i\frac{n}{Q^2}\right)}~, \label{p-integral}
5189: %\end{align}
5190: Carrying out the $p$-integral first, we find~\footnote
5191:   {Here, we are temporarily shifting
5192:   the contour as $\br \, \rightarrow\, \br-i0$
5193:   to avoid the pole $p=0$.
5194:    We eventually restore it back to $\br$ {\sl after}
5195:   taking the difference
5196:   $a_n(p) \equiv a_n^+(p)-a_n^-(p)$.
5197:   The final result \eqref{ImZ final} remains intact,
5198:   even if another contour shift $\br+i0$ is taken, as is
5199:   easily checked.}
5200: \begin{align}
5201:   & \int_{\mathbb{R}-i0}\frac{\rmd p}{\sqrt{2k}}\, a^+_n(p)
5202:   e^{-2\pi t_c \frac{\alpha'}{4} \frac{p^2}{2k}}
5203: = (-1)^{n+1}\frac{i\sqrt{tk}}{\sqrt{2}} \int_{\mathbb{R} + i\xi^+_n} \frac{\rmd
5204: p'}{\sqrt{2k}}\, \frac {e^{\pi \left(\sqrt{\frac{\alpha'}{2}}\frac{p'}{2}-i
5205: \frac{nk}{2}\right) -2\pi t_o \frac{\alpha'}{4}\frac{p^{'2}}{2k}}} {\cosh \pi
5206: \left(\sqrt{\frac{\alpha'}{2}}\frac{p'}{2}-i\frac{nk}{2}\right)}~.
5207: \qquad  \label{evaluation 2}
5208: \end{align}
5209: Finally, we shift the contour back:
5210: $\br+i\xi_n^+ \, \rightarrow\,
5211: \br$ so that  the open string momentum $p'$ is real-valued. In this
5212: step, we cross the poles so need to take care of pole contributions.
5213: (See Figure 1.)
5214: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5215: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5216: 
5217: \begin{figure}[htbp]
5218:     \begin{center}
5219: %    \includegraphics[width=0.5\linewidth,keepaspectratio,clip]
5220:   \includegraphics[width=13cm,height=10cm]
5221:      {contour1.eps}
5222:     \end{center}
5223:     \caption{ Deformation of the contour from the broken line
5224:     to the solid line picks up pole contributions.}
5225:     \label{contour1}
5226: \end{figure}
5227: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5228: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5229: 
5230: 
5231: The relevant poles are located at
5232: \begin{align}
5233:  & \sqrt{\frac{\alpha'}{2}}p' = i {\al_m }~, ~~~ \al_m \equiv nk
5234: -2\left(m+\frac{1}{2}\right) \quad \mbox{where} \quad m=0,1, \ldots,
5235: \left\lb \frac{nk}{2}-\frac{1}{2}\right\rb~, \qquad
5236: \quad\label{poles}
5237: \end{align}
5238: where $\lb ~~ \rb$ denotes the Gauss symbol, and their residues are
5239: evaluated as $(-1)^{n+1}\frac{i}{\pi} e^{\pi t_o \frac{\al_m^2}{2k}}$.
5240: %Their residues are
5241: %evaluated as
5242: %\begin{align}
5243: % && \mbox{\bf Res}_{p'=i\al_m}
5244: %= (-1)^{n+1}\frac{i}{\pi} e^{\pi t \al_m^2}~.
5245: %\label{residue}
5246: %\end{align}
5247: We thus obtain
5248: \begin{align}
5249:    \int_{\mathbb{R}-i0}\sqrt{\frac{\alpha'}{2}}\frac{\rmd p}{\sqrt{2k}}\, a^+_n(p)
5250:    e^{-2\pi t_c \frac{\alpha'}{4} \frac{p^2}{2k}}
5251: &=  (-1)^{n+1}\frac{i\sqrt{kt_o}}{\sqrt{2}} \int_{-\infty}^{\infty}
5252: \sqrt{\frac{\alpha'}{2}}\frac{\rmd p'}{2k}\, \frac {e^{\pi
5253: \left(\sqrt{\frac{\alpha'}{2}}\frac{p'}{2}-i \frac{nk}{2}\right)
5254: -2\pi t_o \frac{\alpha'}{4} \frac{p^{'2}}{2}}} {\cosh \pi
5255: \left(\sqrt{\frac{\alpha'}{2}}\frac{p'}{2}-i\frac{nk}{2}\right)}
5256: \nn &+ 2(-1)^{n+1} \sqrt{t_o} \sum_{m=0}^{\left\lb
5257: \frac{nk}{2}-\frac{1}{2}\right\rb}\, e^{\frac{\pi t_ok}{2} \left[n-\frac{2}{k}\left(m+\frac{1}{2}\right) \right]^2}~.
5258: \label{evaluation 3}
5259: \end{align}
5260: The integral of $a^-_n(p)$ is calculated in a similar
5261: way. This time, we should start with the contour $\br+i \xi^-_n$
5262: with $(-nk- 1) < \sqrt{\frac{\alpha'}{2}}\xi^-_n <
5263: (-nk+1)$ and, after performing the $p$-integral
5264: first, again shift it back to $\br+i\xi^-_n\,\rightarrow\,\br$. The
5265: relevant pole contributions come from $\sqrt{\frac{\alpha'}{2}}p'=-i\al_m$
5266: ($m=0,1,\ldots, \left\lb \frac{nk}{2}-\frac{1}{2}\right\rb$), and
5267: we obtain
5268: %~\footnote
5269: %   {The relative sign change compared to $a^+_n(p)$ integral
5270: %    originates from the orientation of integration
5271: %    contour surrounding each pole.}
5272: \begin{align}
5273: \int_{\mathbb{R}-i0} \sqrt{\frac{\alpha'}{2}} \frac{\rmd p}{\sqrt{2k}}\, a^-_n(p)\,
5274:     e^{-\pi t_c \alpha' \frac{p^2}{2k}}
5275: &=  (-1)^{n+1}\frac{i\sqrt{t_ok}}{\sqrt{k}} \int_{-\infty}^{\infty}
5276: \sqrt{\frac{\alpha'}{2}}\frac{\rmd p'}{\sqrt{2k}}\, \frac {e^{\pi
5277: \left(\sqrt{\frac{\alpha'}{2}}\frac{p'}{2}+i \frac{nk}{2}\right)
5278: -2\pi t_o \frac{\alpha'}{4} \frac{p^{'2}}{2k}}} {\cosh \pi
5279: \left(\sqrt{\frac{\alpha'}{2}}\frac{ p'}{2}+i\frac{nk}{2}\right)}
5280: \nn &- 2(-1)^{n+1} \sqrt{t_o} \sum_{m=0}^{\left\lb
5281: \frac{nk}{2}-\frac{1}{2}\right\rb}\, e^{\frac{\pi t_o k}{2} \left[
5282: n-\frac{2}{k}\left(m+\frac{1}{2}\right) \right]^2}~.
5283: \label{evaluation 4}
5284: \end{align}
5285: Notice that the relative sign change in the pole term compared to
5286: $a^+_n(p)$ integral originates from the orientation
5287: of integration contour surrounding each pole. Therefore,
5288: %from the elementary calculation
5289: %\begin{align}
5290: % &&
5291: %\frac
5292: %{i e^{\pi \left(\frac{p'}{Q}-i \frac{n}{Q^2}\right)}}
5293: %{\cosh \pi \left(\frac{p'}{Q}-i\frac{n}{Q^2}\right)}
5294: %-
5295: %\frac
5296: %{i e^{\pi \left(\frac{p'}{Q}+i \frac{n}{Q^2}\right)}}
5297: %{\cosh \pi \left(\frac{p'}{Q}+i\frac{n}{Q^2}\right)}
5298: %= \frac
5299: %{2 \sin\left(\frac{2\pi n}{Q^2}\right)}
5300: %{\cosh\left(\frac{2\pi p'}{Q}\right)
5301: %+\cos\left(\frac{2\pi n}{Q^2}\right)}~,
5302: %\end{align}
5303: we find
5304: \begin{align}
5305: \int_0^{\infty}\sqrt{\frac{\alpha'}{2}}\frac{\rmd p}{\sqrt{2k}}\, a_n(p)e^{-2\pi t_c \frac{\alpha'}{4}
5306: \frac{p^2}{2k}} &= \frac{1}{2} \int_{-\infty}^{\infty} \sqrt{\frac{\alpha'}{2}}\frac{\rmd p}{\sqrt{2k}}\,
5307: \left(a_n^+(p) - a_n^-(p)\right)e^{-2\pi \alpha' \frac{t_c}{4} \frac{p^2}{2k}} \nn
5308: %
5309: &= (-1)^{n+1}\frac{\sqrt{t_ok}}{\sqrt{2}} \int_{-\infty}^{\infty}\sqrt{\frac{\alpha'}{2}}\frac{\rmd p'}{\sqrt{2k}}\,
5310: \frac{\sin \left(\pi n k \right) e^{-2\pi t_o \frac{\alpha'}{4}
5311: \frac{p^{'2}}{2k}}} {\cosh \left(\pi \sqrt{\frac{\alpha'}{2}} p'\right) +\cos
5312: \left(\pi n k\right)} \nn
5313: %
5314: &+ 2 (-1)^{n+1} \sqrt{t_o}
5315: \sum_{m=0}^{\left\lb \frac{nk}{2}-\frac{1}{2} \right\rb} \,  e^{\frac{\pi t_o k}{2} \left[
5316: n-\frac{2}{k}\left(m+\frac{1}{2}\right) \right]^2}~.
5317: \label{evaluation 5}
5318: \end{align}
5319: In this way,
5320: we derive the desired open string channel expression of
5321: the total radiation rate;
5322: \begin{align}
5323:  \overline{\cN} &= N^2_{\rm NS} \int_0^{\infty} \frac{\rmd t_o}{t_o}\,
5324: \Big( F_{\rm naive} (t_o) + F_{\rm pole} (t_o) \Big)~, \nn
5325:  F_{\rm naive} (t_o) &= \sqrt{\frac{k}{2}}\int_{-\infty}^{\infty} \sqrt{\frac{\alpha'}{2}} \frac{\rmd p'}{\sqrt{2k}}
5326: \sum_{n=1}^{\infty} \, (-1)^{n+1} \frac{\sin \left(\pi nk\right) e^{-\pi t_o \left( \frac{\alpha'}{2} \frac{p^{'2}}{2k} +
5327: \frac{2n^2}{k}\right)}} {\cosh \left(\pi \sqrt{\frac{\alpha'}{2}}
5328: p' \right) + \cos \left(\pi n k\right)} \,
5329: Z_{\cM}^{(o)}(t_o) \nn F_{\rm pole}(t_o) &= 2 \sum_{n=1}^{\infty}
5330: (-1)^{n+1} \sum_{m=0}^{\left\lb \frac{nk}{2}-\frac{1}{2}\right\rb}
5331: \,e^{\pi t_o \left[\frac{2}{k}\left(m+\frac{1}{2}\right)^2- 2n
5332: \left(m+\frac{1}{2}\right)\right]}\, Z_{\cM}^{(o)}(t_o)~. \label{ImZ
5333: final}
5334: \end{align}
5335: The first term in \eqref{ImZ final} coincides with the total radiation
5336: claimed by \cite{Okuyama:2006zr} modulo inessential numerical factor~\footnote
5337:   {It differs slightly from the one
5338:    given in \cite{Okuyama:2006zr} in that we study
5339:    the fermionic string, while \cite{Okuyama:2006zr} studies the bosonic string.}.
5340:    It remains finite as $t_o\,\rightarrow\, + \infty$. The second
5341: term, which the analysis of \cite{Okuyama:2006zr} missed altogether, is of
5342: crucial importance. It is evident that the $m=0$ term is the leading
5343: contribution for each $n$.\footnote{Here we have assumed that $k>1$.} Recalling $Z^{(o)}_{\cM}(t_o)\, \sim \,
5344: e^{\pi t_o \left(1-\frac{1}{2k}\right)}$ asymptotically (up to
5345: pre-exponential power corrections), each $m=0$ term behaves as
5346: \begin{align}
5347:  \sim e^{\pi t_o \left(\frac{1}{2k}- n\right)
5348: + \pi t_o \left(1-\frac{1}{2k}\right)} = e^{\pi t_o (1-n)} \qquad
5349: \mbox{as} \qquad t_o\,\rightarrow\, + \infty~. \label{evaluation m=0
5350: term}
5351: \end{align}
5352: Therefore, we get the leading contribution from the $n=1$ term,
5353: which shows a massless behavior. Hence, we have reproduced
5354: the Hagedorn-growth behavior expected in \cite{Nakayama:2004yx,Sahakyan:2004cq,
5355: Nakayama:2004ge,Nakayama:2005pk}. Notice that all the $n>1$
5356: contributions are massive, and thus are not relevant in the
5357: ultraviolet regime of closed string radiations.
5358: 
5359: 
5360: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5361: 
5362: \subsubsection{Lorentzian cylinder amplitude}\label{sec:8-2-2}
5363: %
5364: In the previous section, we recasted the total emission number
5365: $\overline{\cN}$ of the rolling D0-brane, defined as the sum over
5366: the on-shell states of emitted closed string \eqref{ImZ 0}, in the
5367: open string channel. Now, by the optical theorem and the channel
5368: duality, we ought to be able to obtain $\overline{\cN}$ equally well
5369: from the cylinder amplitude evaluated in the open channel. In this
5370: section, we shall compute explicitly the cylinder amplitude in the
5371: open string channel and show that its imaginary part reproduces
5372: precisely the result \eqref{ImZ final}. This would serve as a
5373: non-trivial check-point of our previous analysis for the consistency
5374: with unitarity and the open-closed channel duality. Notice in
5375: particular that the channel duality is far from being obvious in the
5376: world-sheet in Lorentzian signature. For definiteness, we continue to
5377: focus on the NS sector.
5378: 
5379: 
5380: We start with the cylinder amplitude with Lorentzian
5381: world-sheet~\footnote
5382:   {Here, we stress the importance of
5383:    taking the world-sheet Lorentzian.
5384:    The Fourier transformation from the closed to
5385:    open channel is well-defined only for
5386:    the Lorentzian $\om_L$ in space-time. Accordingly,
5387:     we need to take
5388:    the Lorentzian world-sheet so that the cylinder amplitude
5389:    becomes well-defined.} $Z_{\rm cylinder}$:
5390: %~\footnote
5391: %  {Notice that \eqref{L cylinder closed} has the
5392: %correct normalization
5393: %consistent with \eqref{ImZ 0}. In fact, for each
5394: %transverse mass $M$,
5395: %the modulus integral yields the propagator
5396: %
5397: %\be \int_0^{\infty}\rmd T_L\,q_c^{\frac{1}{2}
5398: %\left(p^2+M^2-(1-i\hat{\ep})^2\om_L^2\right)} = \frac{1}{\pi}
5399: %\frac{i}{p^2+M^2-(1-i\hat{\ep})^2 \om_L^2}. \ee
5400: %
5401: % It contributes to the imaginary part
5402: %as $\frac{1}{2 \om_{p,M}}\left[\delta(\om_L-\om_{p,M})+
5403: %\delta(\om_L+\om_{p,M})\right]$,
5404: %reproducing correctly \eqref{ImZ 0}
5405: %(because $\left|\Psi(p,-\om)\right|^2
5406: %=\left|\Psi(p,\om)\right|^2
5407: %$). }:
5408: \begin{align}
5409: & Z_{\rm cylinder} \nn
5410: &= i \frac{\alpha'}{2} \int_{s_c^{\rm
5411: UV}}^{s_c^{\rm IR}} \rmd s_c \int_{0}^{\infty}\frac{\rmd p}{\sqrt{2k}}
5412: \int_{-\infty}^{\infty} \frac{\rmd \om_L}{\sqrt{2k}} \, \frac {\sinh
5413: \left(\pi \sqrt{\frac{\alpha'}{2}}p\right)}
5414: {\left[\cosh\left(\pi \sqrt{\frac{\alpha'}{2}}
5415: \om_L\right) +\cosh\left(\pi \sqrt{\frac{\alpha'}{2}}p
5416: \right)\right] \sinh(\frac{\pi}{k} \sqrt{\frac{\alpha'}{2}} p)} \nn
5417: %
5418: & \hskip3cm \times
5419: \frac{q_c^{\frac{1}{4}\alpha'(\frac{p^{2}}{2k}-(1-i\hat{\ep})\frac{\om_L^{2}}{2k})}}
5420: {\eta(q_c)^2} \frac{\th_3(q_c)}{\eta(q_c)}
5421:  \cdot Z_{\cM}^{(c)}(q_c) \cdot \eta(q_c)^2
5422:  \frac{\eta(q_c)}{\th_3(q_c)}~.
5423:  \label{zl}
5424: \end{align}
5425: %
5426: Here, we again adopt the $i\ep$-prescription for the Lorentzian
5427: world-sheet, while the $-i \hat{\ep}$-prescription for the
5428: Lorentzian space-time. The integration is well-defined so long as
5429: $2 \hat{\ep} s_c^{\rm UV} > \ep >0$ is retained.
5430: 
5431: 
5432: %Here, $q_c=e^{2\pi i \tau_c}$ is the modulus of the Lorentzian
5433: %cylinder worldsheet, and the Lorentzian worldsheet is regularized by
5434: %adopting $+i \epsilon$-prescription $\tau_c \equiv s_c +i\ep$, with
5435: %$\epsilon >0$. $s_c^{\rm UV}$ and $s_c^{\rm IR}$ are appropriate
5436: %ultraviolet and infrared regularization cut-off's. They will be set
5437: %to zero after performing integrations. We also regularize the
5438: %Lorentzian spacetime with the $-i \hat\epsilon$-prescription
5439: %with $\hat\epsilon >0$. With these prescriptions,
5440: %the integral over $\om_L$ is convergent
5441: %so long as $2 \hat{\ep} s_c^{\rm UV} > \ep >0$ is retained.
5442: 
5443: 
5444: 
5445: The second line in \eqref{zl} combines contributions of the
5446: SL(2)$_k$/U(1), ${\cal M}$, and the world-sheet ghosts. The ghost
5447: contribution $\eta(q_c)^2 \frac{\eta(q_c)}{\th_3(q_c)}$ is seen to
5448: cancel out the contribution of longitudinal oscillators. Thus, the
5449: amplitude simplifies to
5450: %
5451: \begin{align}
5452: & \hskip0.5cm Z_{\rm cylinder} \nn &= i\frac{\alpha'}{2} \int_{s_c^{\rm
5453: UV}}^{s_c^{\rm IR}} \rmd s_c \int_{0}^{\infty} \frac{\rmd p
5454: }{\sqrt{2k}}\int_{-\infty}^{\infty} \frac{\rmd \om_L}{\sqrt{2k}} \, \frac {\sinh
5455: \left(\pi \sqrt{\frac{\alpha'}{2}}p \right)
5456: q_c^{\frac{1}{4}\alpha'(\frac{p^{2}}{2k}-(1-i\hat{\ep})^2\frac{\om_L^{2}}{2k})} \cdot
5457: Z_{\cM}^{(c)}(q_c) } {\left[\cosh\left(\pi \sqrt{\frac{\alpha'}{2}}
5458: \om_L\right) +\cosh\left(\pi \sqrt{\frac{\alpha'}{2}} p
5459: \right)\right] \sinh(\frac{\pi}{k} \sqrt{\frac{\alpha'}{2}} p)}~. \nn
5460: %\nn
5461: %
5462: %&& \hskip4cm \times
5463: %q_c^{\frac{1}{2}(p^{2}-(1-i\hat{\ep})^2\om_L^{2})} \cdot
5464: %Z_{\cM}^{(c)}(q_c)~.
5465: &
5466: \label{L cylinder closed}
5467: \end{align}
5468: 
5469: 
5470: We now modular transform \eqref{L cylinder closed} to the open string
5471: channel.
5472: Define again the open string modulus as $q_o=e^{-2\pi i \tau_o}$,
5473: where $\tau_o = s_o -  i \ep$ and $s_o = 1/s_c$. Using the Fourier
5474: transform identity:
5475: \begin{align}
5476:  & \int_{-\infty}^{\infty}\rmd x\,
5477: \frac{\sin(\pi a x)}{\sinh(\pi x)} e^{-2\pi i kx} = \frac{\sinh (\pi
5478: a)}{\cosh(2\pi k)+ \cosh (\pi a)} ~, \qquad
5479: (\left|\mbox{Im}\,a\right| < 1) \label{FT formula}
5480: \end{align}
5481: we then obtain
5482: %\footnote
5483: %   {Here, we used
5484: %$$ q_c^{\frac{1}{2}\alpha'\left(p^2-(1-i\hat{\ep})^2\om_L^2\right)}
5485: %= s_o \int_{-\infty}^{\infty}\rmd p' \int_{-\infty}^{\infty}\rmd
5486: %\om_L'\, e^{2\pi i \alpha' pp'}e^{2\pi i \alpha' \om_L \om_L'}\,
5487: %q_o^{\frac{1}{2}\alpha'\left(p^{'2}-(1+i\hat{\ep}')^2\om_L^{'2}\right)}~, $$
5488: %which is derivable by the analytic continuation $\om_L = -i \om$, $\om_L'=i\om'$ from
5489: %$$ e^{i\pi \tau \alpha' \left(p^2+\om^2\right)}
5490: %= \frac{i}{\tau} \int_{-\infty}^{\infty}\rmd p'
5491: %\int_{-\infty}^{\infty}\rmd\om' \, e^{2\pi i \alpha' pp'}e^{2\pi i
5492: %\alpha' \om \om'} e^{-i \frac{\pi}{\tau}\alpha'\left(p^{'2}+\om^{'2}\right)}~.
5493: %$$ }
5494: \begin{align}
5495: & \hskip+0.5cm Z_{\rm cylinder} = \nn
5496: %
5497: & \hskip-0.8cm \frac{i\alpha'}{4} \int_{s_o^{\rm UV}}^{s_o^{\rm IR}}
5498: \frac{\rmd s_o}{s_o} \int_{-\infty}^{\infty}\frac{\rmd p'}{\sqrt{2k}}
5499: \int_{-\infty}^{\infty}\rmd \frac{\om_L'}{\sqrt{2k}}\, \frac {\sinh
5500: \left(\pi \sqrt{\frac{\alpha'}{2}} \om'_L \right)
5501: q_o^{\frac{1}{4}\al'(\frac{p^{'2}}{2k}-(1+i\hat{\ep}')^2\frac{\om_L^{'2}}{2k})} \cdot
5502: Z_{\cM}^{(o)}(q_o) } {\left[\cosh\left(\pi \sqrt{\frac{\alpha'}{2}}
5503: \om'_L \right) +\cosh\left(\pi \sqrt{\frac{\alpha'}{2}}p'
5504: \right)\right]\sinh(\frac{\pi}{k} \sqrt{\frac{\alpha'}{2}} \om_L')}.\,\,\,\,\,\,
5505: \nn
5506: %
5507: %&& \hskip5cm \times
5508: %q_o^{\frac{1}{2}(p^{'2}-(1+i\hat{\ep}')^2\om_L^{'2})} \cdot
5509: %Z_{\cM}^{(o)}(q_o)~.
5510: & \label{L cylinder open}
5511: \end{align}
5512: Again
5513: $s_o^{\rm UV} \equiv 1/s_c^{\rm IR}$, $s_o^{\rm IR}\equiv
5514: 1/s_c^{\rm UV}$ are the cut-off's
5515: and the expression \eqref{L cylinder
5516: open} is well-defined so long as
5517: $2 \hat{\ep}' s_o^{\rm UV}
5518: > \ep$.
5519: 
5520: 
5521: \subsubsection{analytic continuation}\label{sec:8-2-3}
5522: We shall now analytically continue both the space-time and the
5523: world-sheet to the Euclidean signature. We have to carefully make the
5524: continuation so that keeping the original amplitude \eqref{L cylinder
5525: open} unchanged (up to cut-off's). As in the previous section, we
5526: should first Wick rotate in space-time $\om'_L\, \rightarrow\,
5527: e^{i(\frac{\pi}{2}-0)} \om'_L$ with $\om'_L = i \om'$ $(\om' \in
5528: \br)$, and then rotate the world-sheet $s_o \, \rightarrow \, - i
5529: t_o$ ($t>0$).
5530: %
5531: %We have to carefully make the
5532: %continuation so that keeping the original amplitude \eqref{L cylinder
5533: %open} unchanged (up to cut-off's). First, Wick rotate the open
5534: %string channel energy as $\om'_L\, \rightarrow\,
5535: %e^{i(\frac{\pi}{2}-0)} \om'_L$ and set $\om'_L = i \om'$. Then, we
5536: %can safely Wick rotate the worldsheet Schwinger parameter as $s_o \,
5537: %\rightarrow \, - i t_o$ ($t>0$). Notice that we will need
5538: %  to perform the Euclidean rotation in {\em
5539: %   opposite} directions
5540: %  for the closed and open string channels due to
5541: %   the difference of the $i\ep$-prescription.
5542: %
5543: %
5544: We shall omit the cutoff's from now on.
5545: %because we are only
5546: %interested in the IR behavior of open channel (UV behavior
5547: %of closed channel).)
5548: We reach the expression
5549: %
5550: \begin{align} Z_{\rm cylinder} = Z_{\rm naive} + Z_{\rm pole}~, \end{align}
5551: %
5552: where the first part is the contribution from naive continuation,
5553: while the second parts originates from the poles passed over by the
5554: rotated contour: $\om'_L\, \rightarrow\, e^{i(\frac{\pi}{2}-0)}
5555: \om'_L$. See Figure 2.
5556: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5557: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5558: \begin{figure}[htbp]
5559:     \begin{center}
5560: %    \includegraphics[width=0.6\linewidth,keepaspectratio,clip]
5561:    \includegraphics[width=13cm,height=10cm]
5562:      {contour2.eps}
5563:     \end{center}
5564:     \caption{ The $\om_L$-integral with Lorentzian contour (broken line)
5565:     and the Euclidean contour (solid line).}
5566:     \label{contour2}
5567: \end{figure}
5568: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5569: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5570: The first part $Z_{\rm naive}$ is given by
5571: %
5572: \begin{align} & \hspace{+0.2cm} Z_{\rm naive} = \nn
5573: %
5574: & \int_{0}^{\infty} \frac{\rmd t_o}{t_o}
5575: \int_{-\infty}^{\infty}\!\!\frac{\rmd p'}{\sqrt{2k}}
5576: \int_{(1-i0)\mathbb{R}}\frac{\rmd \om'}{\sqrt{2k}}\, \frac {-
5577: \frac{1}{4}\alpha' \sin \left(\pi
5578: \sqrt{\frac{\alpha'}{2}}\om' \right)
5579: q_o^{\frac{1}{4}\alpha'(\frac{p^{'2}}{2k}+\frac{\om^{'2}}{2k})} \cdot Z_{\cM}^{(o)}(q_o) }
5580: {\left[\cos\left(\pi \sqrt{\frac{\alpha'}{2}} \om' \right)
5581: +\cosh\left(\pi \sqrt{\frac{\alpha'}{2}} p'\right)\right]
5582: \sin(\frac{\pi}{k}\sqrt{\frac{\alpha'}{2}} \om')}~.  \quad \label{Z naive}
5583: %\nn
5584: %
5585: %&& \hskip5.5cm \times q_o^{\frac{1}{2}(p^{'2}+\om^{'2})} \cdot
5586: %Z_{\cM}^{(o)}(q_o)~.
5587: \end{align}
5588: %
5589: The second part $Z_{\rm pole}$ arises from the poles located at
5590: \begin{align}
5591:  & \sqrt{\frac{\alpha'}{2}}\om'_L =
5592: \left\{
5593: \begin{array}{ll}
5594:  \sqrt{\frac{\alpha'}{2}}|p'| + {i}
5595:  \left(2m+1\right) & ~~ m\in \mathbb{Z}_{\geq 0}  \\
5596:  -\sqrt{\frac{\alpha'}{2}}|p'| + {i}
5597:  \left(2m+1\right) & ~~ m\in \mathbb{Z}_{<0 }
5598: \end{array}
5599: \right.
5600: \end{align}
5601: %\begin{align}
5602: % && \om'_L = p'+iQ\left(m+\frac{1}{2}\right) ~, ~~~
5603: %\left\{
5604: %\begin{array}{ll}
5605: % m\in \bz_{\geq 0}~, & ~~ p' >0 \\
5606: % m\in \bz_{<0}~, & ~~ p'<0
5607: %\end{array}
5608: %\right.
5609: %\end{align}
5610: %and also,
5611: %\begin{align}
5612: %&& \om'_L = -p'+iQ\left(m+\frac{1}{2}\right) ~, ~~~ \left\{
5613: %\begin{array}{ll}
5614: % m\in \bz_{\geq 0}~, & ~~ p' <0 \\
5615: % m\in \bz_{<0}~, & ~~ p'>0
5616: %\end{array}
5617: %\right.
5618: %\end{align}
5619: whose residues are (after taking the open string channel modulus
5620: Euclidean, $q_o=e^{-2\pi t}$)
5621: \begin{align}
5622:  & \frac{i}{2} \sqrt{\frac{2}{\alpha'}}\cdot
5623: \frac{\sqrt{2}}{2 \pi\sqrt{k}} \frac{e^{\pm \frac{i\pi}{k}t(2m+1)\sqrt{\frac{\alpha'}{2}} |p'| -
5624: \pi t \frac{2}{k}\left(m+\frac{1}{2}\right)^2}} {\sinh \frac{\pi}{k} \left(\pm
5625: \sqrt{\frac{\alpha'}{2}} |p'|+i \left(2m+1\right)\right)} ~.
5626: \end{align}
5627: We thus obtain
5628: \begin{align}
5629:  &
5630: Z_{\msc{pole}} = \int_0^{\infty} \frac{\rmd t_o}{t_o}\, \left\lb
5631: \sum_{m\geq 0} \int_{0}^{\infty}\sqrt{\frac{\alpha'}{2}}\frac{\rmd p'}{\sqrt{2k}}
5632:  -\sum_{m<0} \int_{-\infty}^{0}\sqrt{\frac{\alpha'}{2}}\frac{\rmd p'}{\sqrt{2k}}
5633: \right\rb \,  %\frac
5634: {e^{\frac{i\pi}{k} t  (2m+1)\sqrt{\frac{\alpha'}{2}} p' - \pi t \frac{2}{k}
5635: \left(m+\frac{1}{2}\right)^2}}
5636: %{\sinh \pi Q \left(p'+iQ \left(m+\frac{1}{2}\right)\right)}
5637: \nn & \hskip3cm \times 2\pi i \cdot \frac{i\sqrt{2}}{4 \pi\sqrt{k}} \left\lb
5638: \frac{1}{\sinh \frac{\pi}{k}  \left(\sqrt{\frac{\alpha'}{2}} p'+i
5639: \left(2m+1\right)\right)} + (p' \leftrightarrow -p')
5640: \right\rb Z_{\cM}^{(o)}(q_o)\nn
5641: %\hspace{1cm} + \int_0^{\infty} \frac{\rmd t}{t}\, \left\lb
5642: %\sum_{m\geq 0} \int_{-\infty}^{0}\rmd p'
5643: % -\sum_{m<0} \int_{0}^{\infty}\rmd p'
5644: %\right\rb \, 2\pi i \cdot \frac{iQ}{4 \pi} \frac{e^{-i\pi t Q
5645: %(2m+1)p' - \pi t Q^2\left(m+\frac{1}{2}\right)^2}} {\sinh \pi Q
5646: %\left(-p'+iQ \left(m+\frac{1}{2}\right)\right)} \cdot
5647: %Z_{\cM}^{(o)}(q_o) \nn
5648: & \hspace{1cm} = - 2\sqrt{\frac{2}{k}}  \int_0^{\infty} \frac{\rmd t_o}{t_o}\,
5649: \sum_{m=0}^{\infty} \int_{0}^{\infty}\sqrt{\frac{\alpha'}{2}}\frac{\rmd
5650: p'}{\sqrt{2k}}\, \frac{e^{\frac{i\pi}{k} t (2m+1)\sqrt{\frac{\alpha'}{2}} p' - \pi t
5651: \frac{2}{k}\left(m+\frac{1}{2}\right)^2}} {\sinh \frac{\pi}{k}  \left(
5652: \sqrt{\frac{\alpha'}{2}} p'+i \left(2m+1 \right)\right)}
5653: \cdot Z_{\cM}^{(o)}(q_o) ~. \nn & \label{Z pole}
5654: \end{align}
5655: 
5656: We thus obtained manifestly convergent open string channel
5657: expressions \eqref{Z naive}, \eqref{Z pole} for the cylinder amplitude
5658: in Lorentzian signature of the space-time.
5659: 
5660: \subsubsection{optical theorem at work}\label{sec:8-2-4}
5661: With the Lorentzian (in space-time) cylinder amplitude \eqref{Z naive},
5662: \eqref{Z pole} available, we now apply the unitarity and obtain total
5663: emission number $\overline{\cal N}$ via imaginary part of $Z_{\rm
5664: cylinder}$. In the analysis of \cite{Okuyama:2006zr} only the naive contribution
5665: $Z_{\msc{naive}}$ was considered. Taking the imaginary part picks up
5666: infinite poles located at the real $\om'$-axis (the imaginary
5667: $\om'_L$-axis), depicted in Figure 2. Their contributions yield
5668: %
5669: \begin{align}
5670:   & \mbox{Im}\, Z_{\msc{naive}} \nn
5671:   %
5672:   & = - \frac{1}{2}
5673: \int_{0}^{\infty} \frac{\rmd t_o}{t_o} \int_{-\infty}^{\infty} \sqrt{\frac{\alpha'}{2}}\frac{\rmd p'}{\sqrt{2k}} \sum_{\stackrel{n\neq 0}{n\in \bsz}} \, \pi \mbox{sgn}\,(n) \,
5674: \frac{(-1)^n\sqrt{k}}{\pi \sqrt{2}} \frac {\sin\left(\pi n k\right)
5675: e^{-\pi t_o \left( \frac{\alpha'}{2} \frac{p^{'2}}{2k}+\frac{n^2 k}{2}\right)}}
5676: {\cos\left(\pi n k \right) +\cosh \left(\pi \sqrt{\frac{\alpha'}{2}} p' \right)} \cdot Z_{\cM}^{(o)}(q_o) \nn &=
5677: -\sum_{n=1}^{\infty}\int_{0}^{\infty} \frac{\rmd t_o}{t_o}
5678: \int_{-\infty}^{\infty} \sqrt{\frac{\alpha'}{2}} \frac{\rmd p'}{\sqrt{2k}} \, \frac{(-1)^n\sqrt{k}}{\sqrt{2}} \frac
5679: {\sin\left(\pi n k \right) e^{-\pi t_o \left(\frac{\alpha'}{2}
5680: \frac{p^{'2}}{2k}+\frac{n^2k}{2}\right)}} {\cos\left(\pi n k\right)
5681: +\cosh \left(\pi \sqrt{\frac{\alpha'}{2}}p'\right)} \cdot
5682: Z_{\cM}^{(o)}(q_o)~, \label{ImZ 1st}
5683: \end{align}
5684: reproducing the first term in \eqref{ImZ final}.
5685: 
5686: 
5687: We next evaluate the contribution from the pole contribution
5688: $Z_{\msc{pole}}$ \eqref{Z pole}. As is easily seen, taking the
5689: imaginary part just amounts to extending the integration region of
5690: $p'$ in \eqref{Z pole}
5691: %from $(0,\infty)$
5692: to the whole real axis $(-\infty, \infty)$. By closing the
5693: $p'$-contour in the upper half plane, we thus obtain
5694: \begin{align}
5695: & \mbox{Im}\, Z_{\msc{pole}} \nn
5696: %&\equiv& \frac{1}{2i}\left(
5697: %Z_{\msc{pole}} - Z_{\msc{pole}}^*
5698: %\right) \nn
5699: &=
5700:  i\sqrt{\frac{2}{k}} \int_0^{\infty} \frac{\rmd t_o}{t_o}\,
5701: \sum_{m=0}^{\infty} \int_{-\infty}^{\infty}\sqrt{\frac{\alpha'}{2}}\frac{\rmd p'}{\sqrt{2k}}\, \frac{e^{\frac{i\pi}{k} t_o
5702:  (2m+1)\sqrt{\frac{\alpha'}{2}} p' - \pi t_o \frac{2}{k}\left(m+\frac{1}{2}\right)^2}}
5703: {\sinh \frac{\pi}{k}  \left(\sqrt{\frac{\alpha'}{2}} p'+i
5704: \left(2m+1\right)\right)} \cdot Z_{\cM}^{(o)}(q_o) \nn
5705: %&=& iQ \int_0^{\infty} \frac{\rmd t}{t}\,
5706: %\sum_{m=0}^{\infty} \int_{\bsr + iQ \left(m+\frac{1}{2}\right)}\rmd
5707: %p'\, \frac{e^{i\pi t Q (2m+1)p' + \pi t
5708: %Q^2\left(m+\frac{1}{2}\right)^2}} {\sinh \pi Q p'} \cdot
5709: %Z_{\cM}^{(o)}(q_o) \nn
5710: &=
5711:  2\pi i \cdot i\sqrt{\frac{2}{k}}
5712: \int_0^{\infty} \frac{\rmd t_o}{t_o}\,   \sum_{m=0}^{\infty}
5713: \sum_{\stackrel{n> \frac{1}{k}\left(2m+1\right)} {n\in \bsz_{>0}}}
5714: \, \frac{(-1)^n\sqrt{k}}{\pi \sqrt{2}} e^{-\pi t_o n (2m+1)+\pi t_o
5715: \frac{2}{k}\left(m+\frac{1}{2}\right)^2} \cdot Z_{\cM}^{(o)}(q_o)
5716:  \nn
5717: &= -2  \int_0^{\infty} \frac{\rmd t_o}{t_o} \,
5718: \sum_{n=1}^{\infty} \sum_{m=0}^{\left\lb
5719: \frac{nk}{2}-\frac{1}{2}\right\rb} \, (-1)^n e^{-\pi t_o n (2m+1)+\pi
5720: t_o \frac{2}{k}\left(m+\frac{1}{2}\right)^2} \cdot Z_{\cM}^{(o)}(q_o)~.
5721: \label{ImZ pole}
5722: \end{align}
5723: In the last line, we exchanged order of the double summations. The
5724: final result agrees perfectly with the total emission number
5725: $\overline{\cal N}$ in \eqref{ImZ final} evaluated via direct
5726: computation of the transition amplitudes in Euclidean world-sheet.
5727: 
5728: 
5729: 
5730: 
5731: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5732: \subsubsection{imaginary D-instantons: decaying versus rolling}
5733: In the previous sections, we studied spectral observables in causal
5734: processes involving decay of unstable D-brane and rolling of
5735: accelerated D-brane. The main result of this work is that
5736: transformation of the total emission number $\overline{\cal N}$ and
5737: the cylinder amplitude $Z_{\rm cylinder}$ from the closed string
5738: channel to the open string channel require careful analytic
5739: continuation on the world-sheet and that, unlike other results
5740: claimed in the literatures, the analytic continuation we adopt gives
5741: results consistent with the unitarity via the optical theorem
5742: $\overline{\cal N} = \mbox{Im}\, Z_{\rm cylinder}$. 
5743: In particular, we
5744: found that the cylinder amplitude consists in general of two parts
5745: $Z_{\rm cylinder} = Z_{\rm naive} + Z_{\rm pole}$, and the second
5746: part is crucial for ensuring the unitarity through its imaginary
5747: part. While we dealt with decaying or rolling process of the
5748: D-brane, the rules we developed ought to extend to other real-time
5749: processes such as open string and D-brane dynamics in electric field
5750: or plane-wave field background.
5751: 
5752: In this section, we highlights several important steps we noted in
5753: establishing consistency between the channel duality and the
5754: unitarity.
5755: 
5756: 
5757: Throughout this work, the strategy for recasting the closed string
5758: emission spectra in open string channel was to expand the transition
5759: probability ${\cal P}(\omega, {\bf p})$ in power series 
5760: of `imaginary D-instantons' \cite{Maloney:2003ck,Lambert:2003zr,Gaiotto:2003rm}, viz. 
5761: contributions of localized states at time $2 \pi
5762: i \alpha' W(m,n)$ for decaying D-branes and at time $(2 \pi i \sqrt{\frac{k}{2}}) n
5763: \sqrt{\frac{\alpha'}{2}}$ for rolling D-branes, respectively.
5764: %At that stage, we found that the order we perform the integration
5765: %over the off-shell energy $k_0$ and the sum over the imaginary
5766: %D-instantons was quite important. In other words, we obtained the
5767: %result by first performing the $k_0$ integration and then summing
5768: %over the imaginary D-instantons.
5769: %\begin{eqnarray}
5770: % Im\, Z \neq \frac{1}{\pi} \sum_M
5771: %\int_0^{\infty} dp \int_{-\infty}^{\infty} dk_0\,
5772: %\frac{f(p,ik_0)}{k_0^2+p^2+M^2}~. \label{wrong evaluation}
5773: %\end{eqnarray}
5774: %This is because the function $f(p,ik_0)$ could have extra poles on
5775: %the complex $k_0$-plane (this is our case actually), and the
5776: %right-hand side in \eqref{wrong evaluation} picks up their
5777: %contributions in addition to the desired on-shell pole $k_0=
5778: %i\sqrt{p^2+M^2}$.
5779: %(I suspect it would be the reason why OR's
5780: %argument  \cite{OR} is led to a wrong result...)
5781: 
5782: A crucial difference we noted for the rolling D-brane in NS5-brane
5783: background, $k>1$, that weight of the $n$-th imaginary
5784: D-instanton, $a_n(p)$, is a non-trivial function of $p$. We
5785: emphasized above that the momentum dependence came about because
5786: accelerated D-brane rolls in the two-dimensional subspace
5787: $\mathbb{R}_t \times \mathbb{R}_\phi$. Being process dependent, it
5788: could be that, in general, {\em the weights are exponentially
5789: growing functions of momentum, and their Fourier transformations are
5790: not necessarily well-defined.}
5791: %\footnote
5792: %   {One may define the Fourier transformation of such
5793: %    exponentially growing functions as the `distributions' or
5794: %    the `generalized functions'.
5795: %    However, this is essentially the same thing
5796: %   as the contour deformation manipulation given here.}
5797: This was indeed the case for the rolling D-brane case. We thus
5798: prescribed the Fourier transform of the D-instanton weight by
5799: analytic continuation via a deformed integration contour. The
5800: prescription then yielded in the open string channel the
5801: contribution $Z_{\rm pole}$ beyond the naive one $Z_{\rm naive}$.
5802: Moreover, whereas the naive contribution is always ultraviolet
5803: finite, the pole contribution exhibited ultraviolet divergence.
5804: Since $\overline{\cal N}$ (or higher spectral moment) is ultraviolet
5805: divergent, we concluded that the presence of ultraviolet divergent
5806: $Z_{\rm pole}$ is crucial for consistency with the unitarity and the
5807: channel duality.
5808: 
5809: From mathematical viewpoint, we found that the pole contribution
5810: $Z_{\rm pole}$ in \eqref{ImZ final} is present in so far as we adopt
5811: mathematically well-posed prescription of the Fourier transform.
5812: From physics viewpoint, we can also argue that the first term
5813: $Z_{\rm naive}$ by itself cannot be the correct answer and the
5814: second term $Z_{\rm pole}$ ought to dominate over the first one.
5815: 
5816: In the range $k>1$, it is easy to see that 
5817: $\mbox{Im}\,Z_{\rm
5818: naive}$ can take a negative value if we tune the value $k$ suitably
5819: within this range. If $Z_{\rm naive}$ is all there is for the
5820: cylinder amplitude, the negative value of its imaginary part
5821: contradicts with the fact that the total emission number
5822: $\overline{\cal N}$ is positive by definition. Moreover, for
5823:  integral value of $k$, which corresponds to rolling
5824: D-branes in $k$ coincident NS5 backgrounds, we observe that the
5825: first term $Z_{\rm naive}$ vanishes identically since the integrand
5826: vanishes. The above observations indicate that extra contribution
5827: ought to be present to the cylinder amplitude beyond the naive
5828: contribution, $Z_{\rm naive}$.
5829: 
5830: On other other hand, we do not have any contradiction of the
5831: cylinder amplitude with the unitarity once the contribution $Z_{\rm
5832: pole}$ is taken into account. This is because $Z_{\rm pole}$ is
5833: dominant (generically divergent) over $Z_{\rm naive}$ and always
5834: positive. We conclude that our prescription for the cylinder
5835: amplitude renders the total emission number, as extracted from the
5836: optical theorem as $\overline{\cal N} = \mbox{Im}\,Z$ always positive
5837: and well-defined.
5838: 
5839: 
5840: The situation is in sharp contrast to that for decaying D-brane
5841: case. There, as recapitulated in section 2, the D-instanton weights
5842: were constant ($a_{n,m} = 1$), so the issue of Fourier transform was
5843: void from the outset. Again, as explained in section 2, the momentum
5844: independence came about because unstable D-brane decays at rest (or
5845: trivially Lorentz boosted). The situation in NS5-brane phase $k > 1$ is also in contrast to that in extreme string phase 
5846: \cite{Giveon:2005mi},
5847: $1/2 \le k \leq 1$, or in `out-going' radiation in nonextremal
5848: NS5-brane background (which involves two-dimensional black hole
5849: geometry) \cite{Nakayama:2005pk}. For these, 
5850: the leading weight $a_1(p)$ is a bounded
5851: function and have well-defined Fourier transformation. Thus, there
5852: does not arise any extra contribution beyond $Z_{\rm naive}$. We
5853: thus obtain via optical theorem an ultraviolet finite total emission
5854: number.\footnote
5855:   {Even in the deep stringy phase $1/2 \le k \leq 1$, $a_n(p)$ is
5856:   exponentially divergent for sufficiently large $n$.
5857:   Therefore, the formula given in \cite{Okuyama:2006zr} have to be still
5858:   corrected. However, only the $n=1$ term
5859:   could cause the Hagedorn divergence as noted above.
5860: %   (Recall \eqref{evaluation m=0 term}.)
5861:    Hence, this correction does
5862:    not modify ultraviolet behavior of the emission number density.}
5863: 
5864: In the previous work \cite{Nakayama:2005pk}, we also noted that the first
5865: D-instanton weight $a_1(p)$ is identifiable with the `grey body
5866: factor' $\sigma(p)$ in the total emission number $\overline{\cal
5867: N}$. There, the identification was based on saddle-point analysis
5868: valid at large mass $M \rightarrow \infty$ in the closed string
5869: channel. The present result in the open string channel, where the
5870: leading ultraviolet divergence arises from the weight $a_1(p)$, then
5871: supports the identification.\footnote{Footnote 3 of \cite{Okuyama:2006zr} claims
5872: the saddle-point approximation used in our earlier works is invalid.
5873: We disagree with their claim: the relevant 
5874: integral is of the type
5875: %
5876: \begin{align} \int^\infty \rmd p \, \exp \Big[-M f\Big({p \over
5877: M}\Big)\Big]. \nonumber \end{align}
5878: %
5879: As $M \rightarrow \infty$, the saddle point approximation is well
5880: justified in so far as  
5881: $$
5882: f(p_*/M) \sim \cO(1)~, ~~~ f''(p_*/M) >0~, ~~~ 
5883: f^{(2n)}(p_*/M) \sim \cO(1)~, ~~ (n \geq 2)~, ~~~ (p_*~:~
5884: \mbox{saddle})~,
5885: $$
5886: and this is indeed our case.}
5887: 
5888: 
5889: 
5890: \subsubsection{comparisons}
5891: %In recasting the closed string emission spectra in the open string
5892: %channel, it is by now clear through this work why the prescription
5893: %of \cite{KLMS} (which studied the decay of unstable D-brane) gave
5894: %the correct result, while the prescription proposed of \cite{OR}
5895: %(which studied the rolling of accelerated D-brane) did not. In order
5896: %to refine the analysis \cite{KLMS}, we argued that one should begin
5897: %with the Lorentzian worldsheet, and then make the Wick rotation to
5898: %the Euclidean worldsheet carefully. This is the prescription we
5899: %undertook in the present work.
5900: 
5901: 
5902: From our analysis, it became clear the reason why \cite{Karczmarek:2003xm}
5903: obtained the correct result for the decay of unstable D-brane is
5904: because the contour rotation in Fourier transform did not encounter
5905: any pole (since the D-instanton weights $a_n(p)$ were
5906: $p$-independent constants), and the naive manipulation yielded the
5907: correct result. In \cite{Okuyama:2006zr}, the prescription of \cite{Karczmarek:2003xm} was
5908: taken literally also for the rolling of accelerated D-brane. It was
5909: then concluded that $Z_{\rm naive}$ refers to the total cylinder
5910: amplitude. We showed throughout this work that this is incorrect
5911: since it overlooked the pole contribution $Z_{\rm pole}$. After all,
5912: only after taking this extra contribution into account, we showed
5913: that the cylinder amplitude is consistent with the channel duality
5914: and the unitarity.
5915: 
5916: Finally, we find it illuminating to understand why $\overline{\cal
5917: N}$ exhibited Hagedorn divergence in the two-dimensional string
5918: theory studied in \cite{Klebanov:2003km}, whereas it is ultraviolet finite in
5919: the linear dilaton background studied in \cite{Karczmarek:2003xm} in
5920: two-dimensional space-time (that is, $c_{\msc{eff}}=0$). The reason
5921: is because the boundary wave function (D-brane transition amplitude)
5922: of the former has non-trivial $p$-dependence that exponentially
5923: diverges, whereas the latter does not.
5924: 
5925: We finish this subsection with a comment on the origin of the ``black hole - string transition" from the open string viewpoint. Operationally, the transition occurs because in the double summation of \eqref{ImZ pole}, the lightest contribution $(m=0, n=1)$ is outside the rage of summation for $ k<1$. In other words, the lightest open string exchange mode, contributing to the power-like divergence  of the imaginary part is projected out. This suggests that the long range interaction between the brane is drastically different between $k<1$, and $k>1$. It would be of great interest to uncover this phenomenon and explain the intuitive reason for the breaking of the tachyon - radion correspondence from the open string viewpoint.
5926: 
5927: 
5928: 
5929: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5930: 
5931: \subsection{Black hole - string transition}\label{sec:8-3}
5932: It has been a recurrent theme \cite{'tHooft:1987tz,Holzhey:1991bx,Horowitz:1996nw,Susskind:1993ws,Sen:1995in} that an elementary particle
5933: or a string is a black hole: a configuration consisting of
5934: (multiple) strings with high enough total mass is equivalent to a
5935: black hole of the same mass and other conserved charges as we have reviewed in section \ref{sec:4}. This brings
5936: a question whether a given configuration is most effectively
5937: described in terms of strings or black holes. By the black hole - string transition, we will refer to such change of the effective
5938: description for a configuration involving massive string
5939: excitations. Roughly speaking, the string is dual to the black hole
5940: and vice versa.
5941: 
5942: An immediate, interesting question is whether the two-dimensional
5943: black hole geometries is also subject to the black hole - string
5944: transition and if so what precisely the dual of the geometries would
5945: be. In this section, we shall investigate this transition by
5946: studying rolling dynamics of a D0-brane placed on the background. If
5947: the background undergoes the transition between the black hole and
5948: the string configurations, propagation of a probe D0-brane would be
5949: affected accordingly. The transition is triggered by $k$ or
5950: $\kappa$, which measures characteristic curvature scale of the
5951: background measured in sting unit and hence string world-sheet
5952: effects. We shall explore a signal of the transition by examining
5953: spectral distribution of the closed string radiation out of the
5954: rolling D0-brane. Other physical observables associated with
5955: D0-brane would certainly be equally viable probes. Though
5956: straightforward to analyze, in this work, we shall not consider
5957: them.
5958: 
5959: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5960: \subsubsection{probing black hole - string transition via D-brane}\label{sec:8-3-1}
5961: 
5962: In section \ref{sec:8-2}, we observed that $\cN(M)_{\msc{in}} \gg
5963: \cN(M)_{\msc{out}}$ for both the supersymmetric and bosonic string
5964: theories in case the string world-sheet effects are weak enough, viz.
5965: $k > 1$ and $\kappa > 3$, respectively. Obviously, such behavior can be interpreted as indicating that the background on which the
5966: radiative process takes place is indeed a black-hole: D0-brane falls
5967: into the horizon and subsequent radiation is mostly absorbed by the
5968: black hole. On the other hand, the behavior that $\cN
5969: (M)_{\msc{in}}\sim \cN (M)_{\msc{out}} \gg \rho(M)^{-1}$ for $k<1$
5970: or $\kappa < 3$ does not seem to bear features present in the black
5971: hole background: while D0-brane falls inward, subsequent radiation
5972: is not mostly absorbed by the black hole but disperse away. Since
5973: this is the regime where the string world-sheet effects are
5974: significant, the background may be described most effectively in
5975: terms of strings. We are thus led to conclude that the background,
5976: whose stringy effects are controlled by the parameter $k$ or
5977: $\kappa$, would make a phase-transition between the black hole and
5978: the string across $k=1$ or $\kappa = 3$. In a different physical
5979: context, this so-called ``black hole - string transition" was studied
5980: recently \cite{Karczmarek:2004bw,Giveon:2005mi}. What distinguishes our consideration and
5981: result from \cite{Karczmarek:2004bw,Giveon:2005mi} is that we are probing possible
5982: phase-transition of the (closed string) background by introducing a
5983: D0-brane in it and studying open string dynamics.
5984: 
5985: %From the sigma model perspective such effects
5986: %are typically based on
5987: %the world-sheet instanton corrections
5988: %and it is difficult to see even from
5989: %the so called ``exact $\alpha'$ corrected background".
5990: 
5991: Possible existence of such a phase transition was first hinted in
5992: \cite{Kutasov:1990ua} in the closed string sector, where they
5993: observed that the $\cN=2$ Liouville superpotential becomes
5994: normalizable once $k>1$ and it violates the Seiberg bound. Recall
5995: that the marginal interaction term is
5996: \begin{align}
5997: S^{\pm} = \psi^{\mp} e^{-\frac{1}{\cQ}(\phi \pm iY)}~,
5998: ~~~(\cQ=\sqrt{2/k})
5999: \end{align}
6000: for the $\cN=2$ Liouville theory, and
6001: \begin{align}
6002: S^{\pm} =  e^{-\frac{1}{\cQ}(\phi \pm \sqrt{1+\cQ^2}iY)}
6003: \equiv e^{-\sqrt{\frac{\kappa-2}{2}} \phi \mp
6004: \sqrt{\frac{\kappa}{2}} iY}~,
6005: ~~~ (\cQ=\sqrt{2/(\kappa-2)})~,
6006: \end{align}
6007: for the bosonic sine-Liouville theory, respectively. Both
6008: interactions are normalizable (exponentially falling off in the
6009: asymptotic far region) if the curvature is sufficiently small that
6010: $k>1$ or $\kappa
6011: >3$ is satisfied. As is well-known, $\cN=2$ Liouville or
6012: sine-Liouville theory is T-dual to the $SL(2;\br)/U(1)$ coset theory
6013: \cite{FZZ,Giveon:1999px,Giveon:1999tq,Hori:2001ax}, so the condition on the level $k$ or $\kappa$
6014: ought naturally to descend to the two-dimensional black hole
6015: description. Indeed, such aspect was discussed in \cite{Karczmarek:2004bw}
6016: purely in the language of the $SL(2;\br)/U(1)$ coset theory (see
6017: also \cite{Hori:2001ax}). Their reasoning is closely related to the
6018: non-formation of the black hole in two-dimensional string theory
6019: (see also \cite{Friess:2004tq} for the discussion concerning this
6020: issue from the matrix model viewpoint).
6021: %%%
6022: In the strong curvature regime, $k<1$, the background is described
6023: more effectively in terms of the $\cN=2$ Liouville theory as it is
6024: weakly coupled. Evidently, the black hole interpretation of the
6025: $SL(2;\br)/U(1)$ theory is less clear in this region, because the
6026: classical $\cN=2$ Liouville theory does not admit an interpretation
6027: in terms of black hole geometry in any obvious way.
6028: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6029: 
6030: We emphasize that such black hole - string transition is not likely
6031: to arise perturbatively and could arise only from nonperturbative
6032: string world-sheet effects as we have reviewed in section \ref{sec:3} and \ref{sec:4}. For instance, tree-level closed string
6033: amplitudes are manifestly analytic with respect to the level $k$.
6034: These amplitudes exhibit a finite absorption rate (thus displaying
6035: the non-unitarity of the reflection amplitudes) regardless of the
6036: value of $k$. In fact, finite-$k$ correction to the amplitudes yield
6037: an irrelevant phase-factor \cite{Dijkgraaf:1992ba,Giveon:2003wn}.
6038: 
6039: 
6040: %The first sign of the $k = 1$ transition in the closed string
6041: %sector appears as the improved unitarity bound
6042: %and the allowed (normalizable) discrete states,
6043: %which is essentially associated with the one-loop amplitudes.
6044: 
6045: However, as was first observed in \cite{Nakayama:2004ge}, situation changes
6046: drastically if we consider the closed string radiation from the
6047: rolling D-brane in such a background. In \cite{Nakayama:2004ge}, it was shown
6048: that the distribution of radiation off D0-brane in extremal
6049: NS5-brane background becomes ultraviolet finite for $k < 1$. In the
6050: previous section, extending the analysis of \cite{Nakayama:2004ge}, we have
6051: shown that the $k=1$ transition shows up manifestly in the open
6052: string sector in the sense that branching ratio between the incoming
6053: and the outgoing radiation distribution (as well as spectral
6054: moments) behaves very differently across $k=1$.
6055: %as a physical observable in the open string sector.
6056: Remarkably, retaining finite $1/k$-correction, which originated from
6057: consistency with the exact reflection relations, was crucial in
6058: obtaining physically sensible results {\sl even for} $k\gg 1$.
6059: Cancellation between the radiation distribution and the exponential
6060: growth of the density of states at large $M$ is quite nontrivial,
6061: and relied crucially on precise functional dependence on $k$.
6062: %We have seen this cancellation is independent of $k (>1)$.
6063: 
6064: An `order-parameter' of the transition is thus provided by the
6065: radiation distribution of rolling D-brane. The phase transition
6066: across $k=1$ is that while the radiation distribution from the
6067: falling D-brane exhibits powerlike ultraviolet divergence for $k>1$,
6068: it becomes finite for $k<1$. Thus, the rolling D-brane in the $k<1$
6069: regime does {\it not} yield a large back-reaction unlike the $k>1$
6070: case. This is also consistent with the assertion that black hole
6071: cannot be formed in the two-dimensional string theory: It seems
6072: difficult to construct two-dimensional black hole by injecting
6073: D-branes to the linear dilaton (or usual Liouville) theory.\footnote
6074: {Such a possibility was proposed in \cite{Karczmarek:2004bw}.}
6075: %%%%
6076: 
6077: It is also worth mentioning that the radion-tachyon correspondence
6078: is likely to fail in the two-dimensional string theory ($k=1/2$). In
6079: fact, had we have such a correspondence, the rolling radion of the
6080: D0-brane could be identified with the rolling tachyon of the
6081: ZZ-brane in the Liouville theory. On the other hand, it is known
6082: that the radiation distribution of the-ZZ brane exhibits a powerlike
6083: ultraviolet divergence \cite{Klebanov:2003km} at leading order in string
6084: perturbation theory, while that of the falling D0-brane does not.
6085: 
6086: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6087: \subsubsection{holographic viewpoint}\label{sec:8-3-2}
6088: 
6089: The  black hole - string transition across $k=1$ also has a natural
6090: interpretation in terms of the holographic principle, as recently
6091: discussed in \cite{Giveon:2005mi}. Adding $Q_1$ fundamental strings to $k$
6092: NS5-branes, one obtains the familiar bulk geometry of the
6093: $AdS_3/CFT_2$-duality. In this context, the density of states of the
6094: dual conformal field theory is given by the naive Cardy formula
6095: $S=2\pi\sqrt{\frac{cL_0}{6}}+2\pi\sqrt{\frac{\bar{c}\bar{L}_0}{6}}$
6096: with $c = 6 k Q_1$ for $k>1$, but not for $k<1$. Rather, the central
6097: charge that should be used in the Cardy formula is replaced by an
6098: effective one $c_{\rm eff}= 6Q_1(2-\frac{1}{k})$ \cite{Kutasov:1990ua}. The
6099: similar effects also showed up in the double scaling limit of the
6100: `little string theory'(LST) \cite{Giveon:1999px,Giveon:1999tq}.\footnote
6101:    {Even though the original `little string theory' is the theory of
6102:    NS5-brane, so $k$ should be positive integer-valued,
6103: one can also consider models with fractional value of the level $k$,
6104: which is less than 1 generically. This is achieved by considering
6105: the {\em wrapped\/} NS5-brane backgrounds, or compactifications on a
6106: Calabi-Yau threefold having rational singularity \cite{Giveon:1999zm}.  From
6107: the regularized torus partition function, one can prove that there
6108: is no normalizable massless states (corresponding to the
6109: `Lehmann-Symanzik-Zimmerman-poles' \cite{Aharony:2004xn}) in such string vacua
6110: if $k<1$, as was discussed in {\em e.g.} \cite{Eguchi:2004ik,Eguchi:2004yi}. }
6111: %%%%%%%
6112: We shall now show that such change of the central charge is also
6113: imperative for reproducing the closed string radiation distribution
6114: correctly from the dual holographic picture.
6115: 
6116: It is an interesting attempt to reproduce the phase transition in
6117: the radiation distribution of rolling D-brane across
6118: %of the closed string emission rate
6119: $k=1$ from the holographic viewpoint. In \cite{Sahakyan:2004cq}, it was
6120: proposed that the rolling D-brane should correspond to the decay of
6121: a certain defect in the dual LST. We shall now extend that analysis
6122: to the $k<1$ case and explore the phase-transition. The relevant
6123: holographic description is based on the following two assumptions.
6124: \begin{enumerate}
6125: \item \underline{fixed radiation number distribution}: The radiation distribution for a fixed mass $M$ is determined
6126: by large $k$ behavior of the pressure in the far future (past). This
6127: is equivalent to the statement that the decay of the radion is
6128: described by a `holographic tachyon condensation'. We assume that
6129: there is no phase transition at $k=1$ for a fixed mass $M$.\footnote
6130:    {Theoretically, there is no reason to exclude
6131:     a finite $1/k$ correction here.
6132:      We only need this assumption phenomenologically
6133:     in order to reproduce the ten-dimensional calculation
6134:     even for $k>1$. A priori,
6135:      the tachyon condensation (in the critical bosonic string)
6136:     itself may receive large string world-sheet corrections. In the
6137:     Dirac-Born-Infeld action analysis, such potential corrections
6138:     were completely dropped out.}
6139: In our convention, the distribution is given by
6140: \begin{equation}
6141: \cN(M)_{\rm LST} \sim e^{-2\pi M \sqrt{\frac{k}{2}}} \ .
6142: \label{asum}
6143: \end{equation}
6144: 
6145: \item \underline{change of density of states}: The final density of closed little string states
6146: in the `holographic tachyon condensation' is given by the square
6147: root of the full nonperturbative density of states in LST. As is
6148: discussed in \cite{Giveon:2005mi}, the full nonperturbative density of states
6149: of the LST is believed to exhibit a phase transition at $k=1$: for
6150: $k>1$, the density of states is related to the Hawking temperature
6151: %of the geometry
6152: %(Euclidean radius of the asymptotic cigar)
6153: as
6154: \begin{equation}
6155: n(M)_{\rm LST} \sim e^{{4\pi} M \sqrt{\frac{k}{2}}} \ .
6156: \label{asuma}
6157: \end{equation}
6158: %%%
6159: In other words, the Hagedorn temperature in LST should
6160: be equated with the Hawking temperature \cite{Aharony:1998ub}
6161: (see also, {\em e.g.} \cite{Harmark:2000hw,Berkooz:2000mz,Kutasov:2000jp}).
6162: %%%
6163: 
6164: On the other hand, for $k<1$, because of the non-normalizability of
6165: the black hole excitation, the nonperturbative density of states of
6166: the LST is equivalent to the density of states of the (dual)
6167: perturbative string theory \cite{Giveon:2005mi}:
6168: \begin{equation}
6169: n(M)_{\rm LST} \sim e^{4\pi M \sqrt{1-\frac{1}{2k}}} \ .
6170: \label{asumb}
6171: \end{equation}
6172: \end{enumerate}
6173: With these assumptions, we can estimate the average radiation number
6174: of the `holographic tachyon condensation' to be
6175: \begin{align}
6176: \overline{\cN}_{\rm LST} = \int_0^\infty \dd M \cN(M) \,
6177: \sqrt{n(M)_{\rm LST}} \ . \nonumber
6178: \end{align}
6179: Note that, in contrast to the bulk string theory calculation, we
6180: have no integration over the radial momentum. Substituting
6181: \eqref{asum} and \eqref{asuma} or \eqref{asumb} according to the
6182: value of $k$, we obtain
6183: \begin{align}
6184: \overline{\cN}_{\rm LST} \sim \int^\infty \dd M \, e^{-{2\pi}M
6185: {\sqrt{\frac{k}{2}}} +{2\pi}M \sqrt{\frac{k}{2}}} \nonumber
6186: \end{align}
6187: for $k>1$, showing powerlike ultraviolet divergent behavior because
6188: of the complete cancellation in the exponent, and
6189: \begin{align}
6190: \overline{\cN}_{\rm LST} \sim \int^\infty \dd M \, e^{-{2\pi M }
6191: {\sqrt{\frac{k}{2}}}+2\pi M \sqrt{1-\frac{1}{2k}}} \ , \nonumber
6192: \end{align}
6193: for $k<1$, showing exponential suppression in the ultraviolet. It is
6194: easy to see that this holographic dual computation reproduces the
6195: bulk computation presented in section \ref{radiation super} up to a
6196: subleading power dependence \eqref{eq:in}, \eqref{eq:ini}.\footnote
6197:    {The exact determination of the pre-exponential power part
6198:    is beyond the scope of the rough estimate presented here.
6199:     It requires the full computational ability in the LST.}
6200: %The discussion on the bosonic case is completely parallel.
6201: 
6202: It should be noted, however, that the cancellation between the
6203: radiation distribution and the density of states has a different
6204: origin in the dual holographic description as compared to the bulk
6205: side. In the holographic description, the origin of the phase
6206: transition is the nonperturbative density of the states in LST while
6207: the radiation distribution at a fixed mass-level $M$ keeps its
6208: functional form unchanged. On the other hand, in the bulk theory,
6209: origin of the cancellation was that the radiation distribution
6210: changes at $k=1$ due to the disappearance of the non-trivial saddle
6211: point in the integration of the radial momentum $p$, while the
6212: density of states is always given by the same formula. Thus the
6213: agreement between the two descriptions is quite non-trivial and we
6214: believe that our results provide yet another evidence of the
6215: holographic duality for the NS5-brane and black hole physics.
6216: 
6217: Though we presented the dual description based on some assumptions,
6218: we can turn the logic around and regard our results as a support for
6219: such assumptions. In particular the quantum gravity phase transition
6220: at $k=1$ in the dual theory proposed in \cite{Giveon:2005mi} is crucial for
6221: understanding the radiation distribution out of a defect decay in
6222: the dual LST. We thus propose our discussion in this section as a
6223: strong support for black hole - string transition.
6224: 
6225: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6226: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6227: \subsection{Boundary states and radiation in Ramond-Ramond sector}\label{sec:8-4}
6228: 
6229: In the case of fermionic black hole background, the rolling D0-brane
6230: would also radiate off closed string states in the Ramond-Ramond
6231: (R-R) sector. In this section, we shall construct R-R boundary state
6232: of the D0-brane and compute radiation rates. Since the world-sheet
6233: theory corresponds to $\cN =2$ superconformal field theory,
6234: correlation functions of the R-R sector and boundary states are
6235: readily obtainable by performing the standard $\cN=2$ spectral flow.
6236: %  \footnote{Alternative possibility is that
6237: %   we use the fermion number spectral flow, but this
6238: %  seems anomalous and inconsistent with
6239: %  the modular bootstrap at least in the Euclidean signature.}
6240: 
6241: We shall begin with discussion regarding properties of reflection
6242: amplitudes for the R-R sector (see \cite{Giveon:2003wn} in the context of two-dimensional black hole). Recall that the reflection relation was given in the
6243: NS-NS sector as
6244: \begin{align}
6245:  U^{-p}_{\om}(\rho,t)^{\rm NS}= \cR^{\rm NS}(-p,\om)
6246: U^{p}_{\om}(\rho,t)^{\rm NS} \quad \mbox{and} \quad
6247: V^{-p}_{\om}(\rho,t)^{\rm NS}= \cR^{\rm NS *}(-p,\om)
6248: V^{p}_{\om}(\rho,t)^{\rm NS}~, \nonumber \end{align}
6249: %
6250: where the exact reflection amplitude $ \cR^{\rm NS}(-p,\om)$ was
6251: defined by
6252: \begin{align}
6253: \cR^{\rm NS}(p,\om) = \frac{\Gamma(1+\frac{ip}{k})\Gamma(+ip)
6254: \Gamma^2(\frac{1}{2}-i\frac{p+\omega}{2})}
6255: {\Gamma(1-\frac{ip}{k})\Gamma(-ip)
6256: \Gamma^2(\frac{1}{2}+i\frac{p-\omega}{2})} \ . \nonumber
6257: \end{align}
6258: To obtain the reflection relation of the R-R sector, we shall
6259: perform the spectral flow by half unit of the $\cN=2$ $U(1)$
6260: current.
6261: 
6262: In sharp contrast to the $\cN=2$ Liouville theory,
6263: %(namely winding states in $SL(2,\br)/U(1)$),
6264: the reflection amplitude now depends on the spin structure of the
6265: R-R sector.\footnote
6266:   {This is because, in the $\cN=2$ Liouville theory,
6267:    the reflection amplitudes for the momentum modes have
6268:    a symmetry under $\omega \to -\omega$.}
6269: Explicitly, the spectral flow is defined as $\omega \to \omega \pm
6270: i$, where the $+$ sign corresponds to spin ($+,-$) states and $-$
6271: sign corresponds to spin ($-,+$) states (in the
6272: $(\frac{1}{2},\frac{1}{2})$ picture): in the $\rho \to \infty$
6273: limit, they are described by $S^{\pm} e^{-\rho} e^{-ip \rho-i\omega
6274: t}$ and the conformal weight is given by $h =
6275: \frac{p^2-\omega^2+1}{4k} + \frac{1}{8}$.
6276: 
6277: Therefore, for the R-R states with spin ($+,-$), the exact
6278: reflection amplitudes become
6279: \begin{equation}
6280: \cR^{\rm R+}(p,\om) = \frac{\Gamma(1+\frac{ip}{k})\Gamma(+ip)
6281: \Gamma^2(1-i\frac{p+\omega}{2})} {\Gamma(1-\frac{ip}{k})\Gamma(-ip)
6282: \Gamma^2(1+i\frac{p-\omega}{2})} \ . \label{refRp}
6283: \end{equation}
6284: Equivalently, if we take spin ($-,+$) R-R states, the exact
6285: reflection amplitudes become
6286: \begin{equation}
6287: \cR^{\rm R-}(p,\om) =
6288: \frac{\Gamma(1+\frac{ip}{k})\Gamma(+ip)\Gamma^2(-i\frac{p+\omega}{2})}
6289: {\Gamma(1-\frac{ip}{k})\Gamma(-ip)\Gamma^2(+i\frac{p-\omega}{2})} \
6290: . \label{refRm}
6291: \end{equation}
6292: It is important to notice that the latter amplitudes have a second
6293: order zero in the light-cone direction $p = \omega >0$ (recall that
6294: $p>0$ in our convention). Similarly, we could derive the reflection
6295: relation for $(\pm,\pm)$ spin structure, but the resultant
6296: amplitudes are compatible only with the analytic continuation to the
6297: `winding time' (in the interior of the singularity), so we would not
6298: delve into details anymore.
6299: 
6300: Consider next the boundary wave function of the R-R sector. For
6301: definiteness, we shall take the absorbed D0-brane \eqref{falling D0}
6302: %
6303: (We focus on the $t_0=0$ case for simplicity.)
6304: %
6305: \begin{align}
6306: {}_{\msc{absorb}}\!\bra{B,{\rm NS};\rho_0} =
6307: \int_0^{\infty}\frac{\dd p}{2\pi} \int_{-\infty}^{\infty}\frac{\dd
6308: \om}{2\pi}\,
6309:   \Psi_{\msc{absorb:NS}}(\rho_0;p,\om) \,
6310: {}^{\widehat{U}}\!\dbra{p,\om} \nonumber \end{align}
6311: %
6312: where
6313: %
6314: \begin{align} \Psi_{\msc{absorb:NS}}(\rho_0;p,\om) =
6315: \frac{\Gamma(\frac{1}{2}-i\frac{p+\omega}{2})
6316: \Gamma(\frac{1}{2}-i\frac{p-\omega}{2})}{\Gamma(1-ip)}
6317: \Gamma\left(1+\frac{ip}{k}\right) \, \left[ e^{-ip\rho_0} -
6318: \frac{\cosh\left(\pi \frac{p-\om}{2}\right)} {\cosh\left(\pi
6319: \frac{p+\om}{2}\right)} e^{+ip\rho_0} \right]~. \nonumber
6320: \end{align}
6321: The boundary wave functions of the R-R sector are then derived by
6322: applying the ${\cal N}=2$ spectral flow $\omega \to \omega \pm i$:
6323: \begin{align}
6324:  \Psi_{\msc{absorb:R}+}(\rho_0;p,\om)
6325: \frac{\Gamma(-i\frac{p+\omega}{2})
6326: \Gamma(1-i\frac{p-\omega}{2})}{\Gamma(1-ip)}
6327: \Gamma\left(1+\frac{ip}{k}\right) \, \left[ e^{-ip\rho_0} +
6328: \frac{\sinh\left(\pi \frac{p-\om}{2}\right)} {\sinh\left(\pi
6329: \frac{p+\om}{2}\right)} e^{+ip\rho_0} \right]~, \nonumber
6330: \end{align}
6331: and
6332: \begin{align}
6333: \Psi_{\msc{absorb:R}-}(\rho_0;p,\om) =
6334: \frac{\Gamma(1-i\frac{p+\omega}{2})
6335: \Gamma(-i\frac{p-\omega}{2})}{\Gamma(1-ip)}
6336: \Gamma\left(1+\frac{ip}{k}\right) \, \left[ e^{-ip\rho_0} +
6337: \frac{\sinh\left(\pi \frac{p-\om}{2}\right)} {\sinh\left(\pi
6338: \frac{p+\om}{2}\right)} e^{+ip\rho_0} \right]~, \nonumber
6339: \end{align}
6340: for the two opposite spin structures. These boundary wave functions
6341: are of course consistent with the exact reflection amplitudes
6342: \eqref{refRp},\eqref{refRm}.
6343: %Note that if we consider $\omega <0$, the
6344: %boundary wave functions for different spin structures are
6345: %reversed with
6346: %each other: $ \Psi_{\msc{falling:R}\pm}(\rho_0;p,-\om)
6347: %= \Psi_{\msc{falling:R}\mp}(\rho_0;p,+\om)$.
6348: 
6349: From these boundary wave functions, we can deduce some physical
6350: properties of the boundary states in the R-R sector:
6351: \begin{itemize}
6352: \item For $k > {1 \over 2}$, in the saddle point approximation of
6353: the radial momentum integral,
6354: radiation distribution of the R-R sector behaves the same as that of
6355: the NS-NS sector. In particular, the absolute value of the
6356: reflection amplitudes behave in the similar manner. Thus, the
6357: radiation distribution of the R-R sector is the same as that of the
6358: NS-NS sector.
6359: %%%
6360: \item For $k = {1 \over 2}$, viz. the two-dimensional black hole,
6361: considerable differences arise. Both boundary wave function and
6362: reflection amplitudes show singularity (or zero) when we take
6363: particular spin structure. It is not clear what the origin of these
6364: singularities of lightlike on-shell states $p = \omega$ would be. We
6365: note that some related discussions were given in \cite{Giveon:2003wn}.
6366: %%%
6367: \item In the mini-superspace limit $k \to \infty$,
6368: the mass gap in the R-R sector vanishes. Therefore, it is well-posed
6369: to question radiation of the massless R-R states off the R-R charge.
6370: From the boundary states given above, we observe that, assuming $p,
6371: \omega > 0$, there is no lightlike pole in $R+$ state while there is
6372: a pole at $p = + \omega$ in the $R-$ state. It is also interesting
6373: to note that, in the subleading contribution proportional to
6374: $e^{+ip\rho_0}$, the pole from the gamma function is cancelled by
6375: the zero in the $\sinh(\pi\frac{p-\omega}{2})$ factor.
6376: 
6377: A possible interpretation is that, roughly speaking, R-R charge is
6378: localized on the incoming light-cone $p = \omega$.\footnote
6379:         {This is true only in the asymptotic region
6380:        $\rho \to \infty$ since the distribution
6381:    near $\rho = 0$ is further related
6382:    to the basis of Ishibashi states
6383:    used in the expansion. In the case of `absorbed'
6384:    basis, there is no contribution from the past horizon.
6385: In addition, because the reflection amplitude vanishes in the $R-$
6386: sector, an observer at $\rho \to \infty$ do not detect any outgoing
6387: wave.}
6388: %\item Since the spectral flow operator does not have
6389: %  a local action due to the lack of discrete $U(1)$ quantization,
6390: %  the GSO projection does not act properly
6391: %  in the open string sector.
6392: %  Thus the `space time SUSY' is further broken
6393: %  by these boundary states.
6394: 
6395: \end{itemize}
6396: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6397: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6398: \subsection{Back to extremal NS5-brane background}\label{sec:8-5}
6399: 
6400: By tuning off $\mu \rightarrow 0$, we are back to the extremal
6401: NS5-brane background. Roughly speaking, the extremal background is
6402: described by the free linear dilaton theory, but crucial differences
6403: from the non-extremal counterpart studied in this work are the
6404: followings:
6405: \begin{itemize}
6406:  \item We have no reflection relation, and the $p>0$ and $p<0$
6407:  states should be treated as independent states.\footnote
6408:    {In this sense, the arguments given in \cite{Nakayama:2004yx}
6409:     are not completely precise, although the main part of
6410:     physical results, say, the closed string
6411:     radiation rates, are not altered.
6412:     }
6413: %%%
6414:  \item The conformal field theory description is not effective in the
6415:  entire space-time: the string coupling diverges at the location of the
6416:  NS5-brane. We cannot completely trace the classical trajectory of
6417:  the D0-brane \eqref{trajectory D0} without facing strong coupling problem.
6418: \end{itemize}
6419: We thus have to keep it in mind that the validity of the conformal
6420: field theory description of extremal NS5-brane is limited to the
6421: sufficiently weak string coupling region.
6422: 
6423: For the extremal NS5-brane, since the relevant conformal field
6424: theory involves linear dilaton and hence is a free theory, we can
6425: introduce the basis of the Ishibashi states as $\dket{p,\om}$,
6426: $(p,\om \in \br)$ associated with the wave function
6427: $\psi^p_{\om}(\rho,t)\propto e^{-\rho} e^{-ip \rho-i\om t}$. Another
6428: non-trivial difference from the non-extremal case is the volume form
6429: of the space-time. Since we have the linear dilaton $\Phi =
6430: \mbox{const}-\rho$ and a flat metric $G_{ij}=\eta_{ij}$, the
6431: relevant volume form becomes
6432: \begin{align}
6433:  \dd \mbox{Vol}= e^{-2\Phi}\sqrt{G}\dd \rho \dd t = e^{2\rho} \dd \rho \dd t~.
6434: \label{vol linear dilaton}
6435: \end{align}
6436: 
6437: Now, the classical trajectory of D0-brane in the extremal NS5-brane
6438: is given by \cite{Kutasov:2004dj}:
6439: \begin{align}
6440:  2\cosh(t-t_0) e^{\rho} = e^{\rho_0}~.
6441: \label{trajectory 2}
6442: \end{align}
6443: The boundary state describing the D0-brane moving along
6444: \eqref{trajectory 2} ought to have the following form:
6445: \begin{align}
6446:  \bra{B;\rho_0,t_0} = \int_{-\infty}^{\infty}\frac{\dd p}{2\pi}\,
6447: \int_{-\infty}^{\infty} \frac{\dd \om}{2\pi}\, \Psi
6448: (\rho_0,t_0;p,\om) \dbra{p,\om}~. \label{symmetric D0 extremal}
6449: \end{align}
6450: The boundary wave function is evaluated as
6451: \begin{align}
6452:  \Psi(\rho_0,t_0;p,\om) &\sim \int \dd v\, \delta\Big(
6453: 2\cosh(t-t_0)e^{\rho}-e^{\rho_0} \Big)\, e^{-\rho-ip\rho-i\om t} \nn
6454: %&=& \int_{-\infty}^{\infty} dt \int e^{\rho} d(e^{\rho})\,
6455: % \delta\left(
6456: %2\cosh(t-t_0)e^{\rho}-e^{\rho_0}
6457: %\right)\, e^{-\rho-ip\rho-i\om t} \nn
6458: &= \int_{-\infty}^{\infty} \dd t \, e^{-ip\rho_0} e^{-i\om t}
6459: \Big[2 \cosh(t-t_0)\Big]^{ip-1} \nn & =
6460: \frac{1}{2}B\left(\frac{1}{2}-i\frac{p+\om}{2},
6461: \frac{1}{2}-i\frac{p-\om}{2} \right) \, e^{-ip \rho_0-i \om t_0}~.
6462: \nn \end{align}
6463: %
6464: In the last expression, we used the formula \eqref{formula 2}. This is
6465: essentially the calculation given in \cite{Nakayama:2004yx}. Finally, by
6466: restoring the important `world-sheet correction factor'
6467: $\Gamma\left(1+i\frac{p}{k}\right)$,\footnote
6468:   {Since in this case we do not have the reflection relation,
6469:    the inclusion of the factor
6470:    $\Gamma\left(1+i\frac{p}{k}\right)$ may sound less affirmative than
6471:    the nonextremal NS5-brane background.
6472:    We argue that the procedure is actually justified by
6473:    considering the limit from the non-extremal case.}
6474: we obtain the boundary wave function
6475: \begin{align}
6476:  \Psi(\rho_0,t_0;p,\om) = \frac{1}{2} B(\nu_+,\nu_-)
6477: \Gamma\left(1+i\frac{p}{k}\right)\, e^{-ip\rho_0-i\om t_0}~.
6478:  \qquad \mbox{where} \qquad \nu_{\pm}
6479: \equiv \frac{1}{2}- i\frac{p\pm \om}{2}~,\nonumber
6480: \end{align}
6481: This is the extremal counterpart of the `symmetric D0-brane' in the
6482: non-extremal NS5-brane background \eqref{symmetric D0}.
6483: 
6484: We can also consider the `half S-brane' counterpart by
6485: %taking the $t_0\,\rightarrow\, \pm \infty$ limit, or
6486: taking the Hartle-Hawking contours depicted in the Figures
6487: \ref{HH-future} and \ref{HH-past}. Namely, for the `absorbed brane',
6488: we obtain
6489: \begin{align}
6490:  {}_{\msc{absorb}}\bra{B;\rho_0,t_0}
6491: = \left( \int_0^{\infty}\frac{\dd p}{2\pi}
6492: \int_0^{\infty}\frac{\dd \om}{2\pi} + \int_{-\infty}^0\frac{\dd
6493: p}{2\pi} \int_{-\infty}^0\frac{\dd \om}{2\pi}\right)\,
6494: \Psi(\rho_0,t_0;p,\om)\, \dbra{p,\om} ~, 
6495: \label{falling D0 extremal}
6496: \end{align}
6497: and for the `emitted brane',
6498: \begin{align}
6499:  {}_{\msc{emitted}}\bra{B;\rho_0,t_0}
6500: = \left( \int_0^{\infty}\frac{\dd p}{2\pi}
6501: \int^0_{-\infty}\frac{\dd \om}{2\pi} + \int_{-\infty}^0 \frac{\dd
6502: p}{2\pi} \int^{\infty}_0\frac{\dd \om}{2\pi}\right)\,
6503: \Psi(\rho_0,t_0;p,\om)\, \dbra{p,\om} ~. \label{emitted D0 extremal}
6504: \end{align}
6505: They are regarded as the counterparts of \eqref{HH symm D0 1} and
6506: \eqref{HH symm D0 2}.
6507: 
6508: The radiation rates were already evaluated in \cite{Nakayama:2004yx,Sahakyan:2004cq}.\footnote{
6509: In this paper, we scaled energy and momentum differently
6510: from \cite{Nakayama:2004yx}. In light of normalization as in \eqref{on-shell
6511: super}, $\om, p$ in this work should be read as $2 \sqrt{k}$ times
6512: $\om, p$ in \cite{Nakayama:2004yx}.} Crucial differences from the non-extremal
6513: case are the followings: We have the `forward radiations' ({\em
6514: e.g.}, the incoming radiation for the absorbed D-brane \eqref{falling
6515: D0 extremal}) only and no `backward radiations' ({\em e.g.}, the
6516: outgoing radiation for the absorbed D-brane). This is because there
6517: is no reflection relation in the extremal case. The forward
6518: radiations behave in the completely same way as the non-extremal
6519: case (that is, in a fermionic two-dimensional black hole with $k >1
6520: $), giving rise to the Hagedorn-like ultraviolet divergence again.
6521: At fixed but large $M$ before integrating over $p$, the partial radiation number distribution
6522: takes again exactly the same asymptotic form as in \eqref{grey body}
6523: except that now the coefficient $2 \pi \sqrt{2k}$ is {\sl not}
6524: interpretable as the inverse Hawking temperature of the black hole.\footnote
6525: {An obvious alternative interpretation could be that, even
6526: for extremal background, the falling D0-brane excites the NS5-brane
6527: above the extremality.} Again, this has to do with the peculiarity
6528: that the Hawking temperature of the two-dimensional black hole is
6529: set by the level $k$, not by the nonextremality $\mu$. On the other
6530: hand, the absence of the backward radiation matches with the
6531: extremality of the background; there is no Hawking radiation.
6532: 
6533: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6534: 
6535: \subsection{More on physical interpretations :
6536: Hartle-Hawking states}\label{sec:8-6}
6537: 
6538: We shall now revisit the boundary states we constructed in this work
6539: and elaborate further on their physical interpretations with
6540: particular emphasis on analogy with the rolling tachyon problem via
6541: the radion-tachyon correspondence. We also elaborate on the fate of
6542: R-R charge carried by the D0-brane. To be concrete, we shall focus
6543: on the cases $k \geq 2$ admitting interpretation in terms of near
6544: horizon geometry of black NS5 branes.
6545: 
6546: The boundary state \eqref{falling D0} describes the late-time rolling
6547: ($t \gg t_0$) of the D0-brane rolling into the black NS5 branes.
6548: %in $t> t_0$.
6549: The relevant D0-brane has the initial condition $\rho=\rho_0$,
6550: $\frac{d\rho}{dt}=0$ at $t=t_0$ and starts to roll down toward the
6551: black hole. After sufficiently long coordinate time elapsed, the
6552: D0-brane gets close to the future horizon (${\cal H}^+$). As
6553: examined in section 4, almost all energy of the D0-brane is absorbed
6554: by the black hole in the form of incoming radiation.
6555: %, which effectively causes a Hagedorn-like divergence.
6556: The incoming radiation is dominated by very massive, and hence
6557: highly non-relativistic closed string excitations. Via the
6558: radion-tachyon correspondence, these states are identifiable with
6559: the `tachyon matter' in the rolling tachyon problem in flat
6560: space-time. On the other hand, we have seen that a small part of
6561: energy escapes to the spatial infinity (${\cal I}^+$) as the
6562: outgoing radiation. We have seen that the spectral distribution is
6563: characterized by the Hawking temperature, and is necessarily
6564: dominated by light modes. This interpretation is quite natural from
6565: the viewpoint of the radion-tachyon correspondence for the extremal
6566: NS5-brane background \cite{Kutasov:2004dj}. Since we are now working with
6567: the non-extremal NS5-brane background, our analysis may be
6568: considered as an evidence that the correspondence is valid even at
6569: finite temperature.
6570: 
6571: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6572: 
6573: What about evolution in the far past $t < t_0$? Here, we face a
6574: subtlety. Recall that the boundary condition defining \eqref{falling
6575: D0} does not allow contributions from the past horizon (${\cal
6576: H}^-$), namely, the basis of Ishibashi states $\dket{p,\om}^U$
6577: %${}^{\widehat{U}}\!\dbra{p,\om}$
6578: does not reproduce the past half of the classical trajectory
6579: \eqref{trajectory D0}. Rather, the NS-NS sector of the D0-brane
6580: boundary wave function appears widely distributed in the space-time
6581: in the far past. This may be interpreted as radiations imploding to
6582: $\rho = \rho_0$ from spatial infinity, but then it is subtle to
6583: trace the R-R charge carried by the D0-brane, created out of the
6584: imploding radiation. Classically, the D0-brane charge density ought
6585: to be localized along the classical trajectory \eqref{trajectory D0}
6586: and hence emanates from the past horizon. Once stringy effects are
6587: taken into account, the charge appears to originate from asymptotic
6588: infinity along the light-cone coordinate. Complete understanding of
6589: this curious feature is highly desirable but we shall relegate it to
6590: future study. Here, instead, we present a simple prescription of
6591: avoiding this subtlety: a version of `Hartle-Hawking' boundary
6592: condition.
6593: 
6594: 
6595: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6596: 
6597: We shall first focus on the absorbed D0-brane boundary state
6598: \eqref{falling D0}. Formally, by construction, we can regard
6599: %(in a formal sense, at least)
6600: the boundary wave function specified by the time-integration over
6601: the `real contour' $\cC= \br$ as in \eqref{Wick rotation 0}. Now, let
6602: us discuss what happens if we choose the `Hartle-Hawking' type
6603: contour instead of the real contour, which connect the Euclidean
6604: time with the future or past half of real time axis at $t=t_0$:
6605: %in order to only describe the future ($t>t_0$) or
6606: %past $(t<t_0)$ half of trajectory \eqref{trajectory D0}
6607: \begin{align}
6608: \cC_{\msc{future}}^{\pm}
6609: = \left(t_0+i\br_{\mp}\right) \cup \left(t_0+\br_{+}\right)~,~~~
6610: \cC_{\msc{past}}^{\pm} = \left(t_0+i\br_{\mp}\right) \cup
6611: \left(t_0+\br_{-}\right)~.
6612: \end{align}
6613: More precisely, we should avoid suitably the branch cuts on
6614: $t_0+i\br$ to render the integral convergent. See Figures
6615: \ref{HH-future} and \ref{HH-past} for details. The superscript $+$
6616: $(-)$ is associated with the positive (negative) energy sector. Note
6617: that the phase-factor $e^{-i\om t}$ behaves well on the lower
6618: (upper) half of complex $t$-plane if $\om $ is positive (negative).
6619: Let us pick up $\cC_{\msc{future}}$. Following the traditional
6620: interpretation of the Hartle-Hawking type wave function, we may
6621: suppose that both the D0-brane and black NS5-brane are created from
6622: `nothing' at $t=t_0$, and then the D0-brane starts to fall down
6623: toward the future horizon along the classical trajectory
6624: \eqref{trajectory D0}. In this prescription, the subtlety we mentioned
6625: above is completely circumvented.
6626: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6627: 
6628: \begin{figure}[htbp]
6629:     \begin{center}
6630:    \includegraphics[width=0.5\linewidth,keepaspectratio,clip]{cont1.eps}
6631:     \end{center}
6632:     \caption{`future Hartle-Hawking contour' : the red (green broken)
6633:         line is the contour $\cC^+_{\msc{future}}$ for $\om > 0$
6634:         ($\cC^-_{\msc{future}}$ for $\om <0$). The `$L$' (`$R$') contour
6635:          should be used if calculating the overlap with
6636:         $L^p_{\om}(\rho,t)$ ($R^p_{\om}(\rho,t)$) for
6637:          the convergence of integral.}
6638:     \label{HH-future}
6639: \end{figure}
6640: 
6641: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6642: \begin{figure}[htbp]
6643:    \begin{center}
6644:     \includegraphics[width=0.5\linewidth,keepaspectratio,clip]{cont2.eps}
6645:     \end{center}
6646:     \caption{`past Hartle-Hawking contour' : the red (green broken)
6647:         line is the contour $\cC^+_{\msc{past}}$ for $\om > 0$
6648:          ($\cC^-_{\msc{past}}$ for $\om <0$).}
6649:     \label{HH-past}
6650: \end{figure}
6651: 
6652: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6653: One may paraphrase the prescription as follows: choosing the
6654: Hartle-Hawking contour $\cC_{\msc{future}}$, we explicitly obtain
6655: \begin{align}
6656: & \hspace{-5mm} {}_{HH +,\,\msc{absorb}}\!\bra{B;\rho_0,t_0} \cr 
6657: &=
6658: \int_0^{\infty}\frac{\dd p}{2\pi}\, \left[
6659: \int_{0}^{\infty}\frac{\dd \om}{2\pi}\,
6660: \Psi_{\msc{symm}}(\rho_0,t_0;p,\om) + \int_{-\infty}^{0}\frac{\dd
6661: \om}{2\pi}\, \cR(p,\om)\Psi^*_{\msc{symm}}(\rho_0,-t_0;p,\om)
6662: \right] \, {}^{\widehat{U}}\!\dbra{p,\om}~, \cr \label{HH falling
6663: D0}
6664: \end{align}
6665: where $\Psi_{\msc{symm}}(\rho_0,t_0;p,\om)$ is defined in
6666: \eqref{symmetric D0}. In fact, by taking $\cC_{\msc{future}}$, we are
6667: only left with the $L^p_{\om}$ ($R^p_{\om}$)-part of the one-point
6668: function for the $\om>0$ ($\om<0$) sector. See figure
6669: \ref{HH-future}. This boundary wave function is formally regarded as
6670: the limit of \eqref{falling D0} under $t_0\,\rightarrow\, -\infty$,
6671: $\rho_0\,\rightarrow\,+\infty$ while keeping $|\rho_0|/|t_0|$
6672: finite. Note that the second (first) term $\propto e^{ip\rho_0-i\om
6673: t_0}$ ($\propto e^{-ip\rho_0-i\om t_0}$) in \eqref{falling D0}
6674: oscillates very rapidly in this limit for $\om >0$ ($\om <0$) and
6675: hence drops off.\footnote
6676:    {More precise argument would be as follows:
6677:     The disk amplitude for a wave packet {\em e.g.}
6678:     $\int \frac{\dd p}{2\pi} \int \frac{\dd \om}{2\pi}\, f(p,\om) \ket{L^p_{\om}}$ is evaluated
6679:     as $
6680:     \lim_{\rho_0\,\rightarrow\,+\infty , \,
6681:     t_0\,\rightarrow\,-\infty}\,
6682:     \int \frac{\dd p}{2\pi} \int \frac{\dd\om}{2\pi} f(p,\om) \Psi(\rho_0,t_0;p,\om)$.
6683:     Then, the rapidly oscillating term in the boundary wave
6684:      function $\Psi(\rho_0,t_0;p,\om)$ cannot contribute for any
6685:      $L^2$-normalizable wave packet $f(p,\om)$ due to the Riemann-Lebesgue theorem.
6686: } The limit just means that the D0-brane moving along the
6687: trajectory \eqref{trajectory D0} is coming from the past infinity
6688: $({\cal I}^-)$, and falling into the future horizon (${\cal H}^+$).
6689: Everything is supposed to be localized over the classical trajectory
6690: in this case.
6691: 
6692: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6693: 
6694: Adopting the past Hartle-Hawking contour $\cC_{\msc{past}}$ for the
6695: boundary state of emitted D0-brane \eqref{emitted D0} is completely
6696: parallel. We take the time-reversal of the above:
6697: \begin{align}
6698: & {}_{HH -,\,\msc{emit}}\!\bra{B;\rho_0,t_0} \cr &=
6699: \int_0^{\infty}\frac{\dd p}{2\pi} \left[ \int_{0}^{\infty}\frac{\dd
6700: \om}{2\pi}\, \Psi^*_{\msc{symm}}(\rho_0,-t_0;p,\om) +
6701: \int_{-\infty}^{0}\frac{\dd \om}{2\pi}\,
6702: \cR^*(p,\om)\Psi_{\msc{symm}}(\rho_0,t_0;p,\om) \right]
6703:  {}^{\widehat{V}}\!\dbra{p,\om}~, \nn \label{HH emitted D0}
6704: \end{align}
6705: which is regarded as the $t_0\,\rightarrow\,+\infty$,
6706: $\rho_0\,\rightarrow\,+\infty$ limit of \eqref{emitted D0}. It
6707: describes the trajectory of D0-brane emitted from the past horizon
6708: ${\cal H}^-$ and escaping to the future infinity ${\cal I}^+$.
6709: 
6710: 
6711: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6712: 
6713: Let us turn to the `symmetric' D0-brane \eqref{symmetric D0}. Naively,
6714: it appears that the prescription is that
6715: \begin{align}
6716: & {}_{HH +,\,\msc{symm}}\!\bra{B;\rho_0,t_0}' \cr &=
6717: \int_0^{\infty}\frac{\dd p}{2\pi} \left[ \int_{0}^{\infty}\frac{\dd
6718: \om}{2\pi}\, 2 \Psi_{\msc{symm}}(\rho_0,t_0;p,\om) \,
6719: {}^L\!\dbra{p,\om} +\int_{-\infty}^{0}\frac{d\om}{2\pi}\, 2
6720: \Psi^*_{\msc{symm}}(\rho_0,-t_0;p,\om) {}^R\!\dbra{p,\om} \right]
6721: \nn \label{HH symm D0 1}
6722: \end{align}
6723: for the future Hartle-Hawking contour $\cC_{\msc{future}}$, and
6724: \begin{align}
6725: & {}_{HH -,\,\msc{symm}}\!\bra{B;\rho_0,t_0}' \cr &=
6726: \int_0^{\infty}\frac{\dd p}{2\pi} \left[ \int_{0}^{\infty}\frac{\dd
6727: \om}{2\pi}\, 2 \Psi^*_{\msc{symm}}(\rho_0,-t_0;p,\om)
6728: {}^R\!\dbra{p,\om} +\int_{-\infty}^{0}\frac{d\om}{2\pi} 2
6729: \Psi_{\msc{symm}}(\rho_0,t_0;p,\om) {}^L\!\dbra{p,\om} \right] \nn
6730: \label{HH symm D0 2}
6731: \end{align}
6732: for the past Hartle-Hawking contour $\cC_{\msc{past}}$.
6733: %For instance, taking $\cC_{\msc{future}}$,
6734: %we are only left the $L^p_{\om}$ ($R^p_{\om}$)-part
6735: %of the one-point function for the $\om>0$ ($\om<0$) sector.
6736: However, this cannot be the whole story. The existence of Euclidean
6737: part of the Hartle-Hawking path-integral enforces the boundary
6738: states to be expanded by the basis smoothly connected to the
6739: Euclidean ones, while $\ket{L^p_{\om}}$, $\ket{R^p_{\om}}$ do not
6740: possess such a property. Consequently, to achieve the correct
6741: Hartle-Hawking states, we ought to make further the projection to
6742: $\cH^U$, ($\widehat{\cH^U}$) for the contour $\cC_{\msc{future}}$,
6743: and to $\cH^V$, ($\widehat{\cH^V}$) for $\cC_{\msc{past}}$. We thus
6744: obtain as the correct Hartle-Hawking states:
6745: \begin{align}
6746: & {}_{HH +,\,\msc{symm}}\!\bra{B;\rho_0,t_0} =  {}_{HH
6747: +,\,\msc{symm}}\!\bra{B;\rho_0,t_0}' \widehat{P_U} \equiv {}_{HH
6748: +,\,\msc{absorb}}\!\bra{B;\rho_0,t_0} ~, \nn & {}_{HH
6749: -,\,\msc{symm}}\!\bra{B;\rho_0,t_0} =  {}_{HH
6750: -,\,\msc{symm}}\!\bra{B;\rho_0,t_0}' \widehat{P_V} \equiv {}_{HH
6751: -,\,\msc{emitted}}\!\bra{B;\rho_0,t_0} ~, \label{relation HH}
6752: \end{align}
6753: where the right-hand sides are already given in \eqref{HH falling D0},
6754: \eqref{HH emitted D0}.
6755: 
6756: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6757: 
6758: 
6759: Remarkably, this feature resembles much that of the S-branes
6760: discussed in \cite{Lambert:2003zr}. Namely, it was shown there that
6761: \begin{equation}
6762: \mbox{half S-brane} ~ \cong ~ \mbox{full S-brane with the
6763: Hartle-Hawking contour} ~. \label{half full S}
6764: \end{equation}
6765: In our case, \eqref{symmetric D0} corresponds to the full S-brane,
6766: while the Hartle-Hawking state \eqref{HH falling D0} (\eqref{HH emitted
6767: D0}) is identifiable as the analogue of the half S-brane describing
6768: unstable D-brane decay (creation) process. The equalities
6769: \eqref{relation HH} suggest that we have roughly identical relation to
6770: \eqref{half full S}.
6771: 
6772: Notice that the parameters $\rho_0$, $t_0$ appear just as phase
6773: factors of boundary wave functions in \eqref{HH falling D0}, \eqref{HH
6774: emitted D0} contrary to \eqref{falling D0}, \eqref{emitted D0}. Namely,
6775: the choice of parameters $\rho_0$, $t_0$ does not cause any physical
6776: difference for the Hartle-Hawking type states : They all can be
6777: regarded as describing the D0-brane moving from ${\cal I}^-$ to
6778: ${\cal H}^+$ (from ${\cal H}^-$ to ${\cal I}^+$) for \eqref{HH falling
6779: D0} (for \eqref{HH emitted D0}) irrespective of $\rho_0$, $t_0$. These
6780: two parameters merely parameterize displacing the trajectory in
6781: two-dimensional black hole background. Similar feature comes about
6782: for the full S-brane with Hartle-Hawking contour as well: It is
6783: equivalent to the half S-brane not depending on any shift of the
6784: origin (the point connecting the real and imaginary times).
6785: 
6786: 
6787: Finally, we remark a comment from the viewpoints of boundary
6788: conformal field theory: in contrast to the original ones
6789: \eqref{falling D0}, \eqref{emitted D0} and \eqref{symmetric D0}, the
6790: Hartle-Hawking boundary states \eqref{relation HH} (or equivalently
6791: \eqref{HH falling D0}, \eqref{HH emitted D0}) are not compatible with
6792: the reflection relations. One may regard the boundary states
6793: \eqref{falling D0} and \eqref{emitted D0} as the `completions' of the
6794: Hartle-Hawking states \eqref{relation HH} so that they satisfy the
6795: reflection relations.
6796: 
6797: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6798: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6799: 
6800: \newpage
6801: \sectiono{Conclusion and Discussions}\label{sec:9}
6802: 
6803: 
6804: In this thesis, we have examined the exact boundary states describing the rolling D-brane in the two-dimensional black hole system. In this final section, we would like to summarize our main results and discuss their physical relevance.
6805: 
6806: In the introduction, we asked three fundamental questions about the nature 
6807: of the quantum gravity, or string theory as a candidate for the theory of 
6808: everything:
6809: \begin{itemize}
6810:  \item Small charge black hole v.s. large charge black hole.
6811:  \item Analyticity v.s. non-analyticity in physical amplitudes.
6812:  \item Unitarity v.s. open closed duality.
6813: \end{itemize}
6814: It would be natural to conclude this thesis by asking how far we can answer 
6815: these questions after our studies on the rolling D-brane in two-dimensional 
6816: black hole system. 
6817: To answer these three fundamental questions, in this paper, we have constructed
6818: the exact boundary states describing the rolling D-brane in the two-dimensional black hole system (section \ref{sec:7}) to probe the quantum geometry. Our  main results are
6819: \begin{itemize}
6820: 	\item The tachyon - radion correspondence is proved for $k>1$ by studying the closed string radiation rate from the $\alpha'$ exact rolling D-brane solution (section \ref{sec:8-1}). 
6821: 	\item The black hole - string transition is observed at $k=1$ in the closed string radiation rate as a physical order parameter (section \ref{sec:8-3}). 
6822: 	\item The consistency between the unitarity and open closed duality is shown to be recovered after a careful treatment of the Wick rotation (section \ref{sec:8-2}).
6823: \end{itemize}
6824:  Although our model is rather a specific one, we can naturally extend our results to draw many universal features of the quantum gravity. In the rest of this section, we recapitulate our arguments and present some discussions with possible future directions to pursue.
6825: 
6826: First of all, we have shown in section \ref{sec:8} that the total emission rate of the rolling D-brane into the two-dimensional black hole system behaves exactly same as that for the rolling tachyon in flat Minkowski space studied in section \ref{sec:5}. This result strongly suggests universal features of the physics associated with the D-brane decay. 
6827: 
6828: The universality is an important concept in any physical system. In our decaying D-brane system, we have shown that the closed string radiation rate is independent of the free parameter $k$ representing the level of the current algebra. Since $1/k$ correction governs the $\alpha'$ correction to the geometry, the physical quantity observed in the decaying D-brane process is independent of the stringy corrections. The classical tachyon - radion correspondence still holds even after introducing the stringy corrections independent of its strength (as long as $k>1$).
6829: 
6830: Indeed, the universal behavior of the closed string radiation should be true from the following simple argument. The D-brane energy that should be released during the decay is always proportional to $1/g_s$, so in the perturbative string computation, we expect a divergence in the radiation rate: otherwise we have to face the missing energy problem.\footnote{At first sight, this viewpoint contradicts our computation that the higher dimensional D-brane shows a power-like {\it finite} emission rate (energy), but this is an artefact of the one-particle decay.} 
6831: 
6832: Furthermore the universality holds under almost every exactly solvable deformations of the model such as an inclusion of the time-like linear dilaton, electric field etc (see section \ref{sec:5},\ref{sec:8}). In \cite{Israel:2006ip}, a similar computation has been performed in the $AdS_3$ space, supporting the universality of the decaying D-brane systems in yet another solvable background. We would like to emphasize again that our results do not depend on the  level $k$ which governs the strength of the world-sheet $\alpha'$ corrections {\it as long as} $k>1.$ This indeed provides a strong support for the tachyon-radion correspondence even at the quantum level.
6833: 
6834: At this point, it is worthwhile mentioning Sen's open-string completeness conjecture \cite{Sen:2003iv,Sen:2003mv,Sen:2004nf}: {\it There is a quantum open string theory (OSFT) that describes the full dynamics of an unstable Dp-brane without an explicit coupling to closed strings. Furthermore, Ehrenfest theorem holds in the weakly coupled OSFT; the classical results correctly describes the time evolution of the quantum expectation values}. The tachyon-radion correspondence directly results in the same conjecture for the rolling D-brane system. The smeared trajectory (or ``moss" around the rolling D-brane) we observed in our rolling D-brane system with exact $\alpha'$ correction is an interesting twist to this conjecture.
6835: 
6836: 
6837: 
6838: 
6839: Secondly, we have shown that the interplay between the analyticity and 
6840: non-analyticity of the physical amplitudes are crucial to discuss the black 
6841: hole - string phase transition. The integration over the radial momenta, 
6842: which at the same time is crucial to prove the tachyon - radion 
6843: correspondence, introduces the non-analyticity in the physical observables, 
6844: resulting in the phase transition.
6845: 
6846:  More precisely, we have directly observed the black hole - string phase 
6847: transition from the exact boundary states for the probe rolling D-brane in 
6848: the two-dimensional black hole background. The phase transition occurs 
6849: exactly when the Hawking temperature of the two-dimensional black hole 
6850: coincides with the Hagedorn temperature of the string background as we 
6851: decrease the charge of the two-dimensional black hole (level of the current 
6852: algebra). Below the phase transition point, the physical interpretations of 
6853: the $SL(2;\br)/U(1)$ coset model as a black hole geometry break down and 
6854: become obscure. Our results show that the tachyon - radion correspondence 
6855: fails at the phase transition point, and the physics associated with the 
6856: D-brane decay changes drastically.
6857: 
6858: Indeed, as we have shown in section \ref{sec:8}, a drastic change occurs when we study the dynamics of rolling D-branes in the two-dimensional black hole with $k<1$. From the arguments given in section \ref{sec:4}, we expect ``black hole - string transition". This transition is subtle even from the exact CFT analysis because in deriving every formulae in the closed string scattering amplitudes, we assume an analyticity in $k$. How can we probe the ``black hole - string transition" with respect to $k$ when the amplitude is analytic in $k$? In the open string channel, $k$ dependence is also analytic in the amplitudes as well. However, if we compute the physical quantities such as closed string emission rate, the non-analyticity with respect to $k$ emerges. In this way, we have succeeded in probing the ``black hole - string transition", as an emergent phase transition, by studying the rolling D-brane dynamics in the background.
6859: 
6860: 
6861: 
6862: It would be interesting to note that not every D-brane can probe the ``black hole - string transition". As we have seen in section \eqref{sec:5-2-5}, the decay of the unstable D0-brane in the Euclidean two-dimensional black hole does {\it not} show any ``black hole - string transition" at $k=1$. The decay rate of unstable D-branes shows a universal property irrespective of the value of $k$. We do not have a good physical explanation of this phenomenon at this moment, but it would be interesting to give a further study and determine which objects can probe the phase transition.
6863: 
6864: 
6865: In the Euclidean signature target-space theory after the analytic 
6866: continuation, the phase transition is induced from the non-perturbative 
6867: $\alpha'$ corrections related to the winding-tachyon condensation. In the 
6868: original Lorentzian signature target-space, one might understand it as a 
6869: thermal (winding) tachyon condensation, or in the real time picture at the 
6870: phase transition point, we would encounter associated (local) Hagedorn 
6871: divergence of the black hole thermodynamics.
6872: 
6873: It is natural to expect that our results on the string - black hole phase 
6874: transition is rather robust and universal. Indeed, the transition is barely 
6875: affected under various marginal deformations of the solvable model such as 
6876: incorporation of the linear dilaton or the electric field. It would be 
6877: interesting to extend our analysis to more realistic higher dimensional 
6878: black hole systems realized in superstring theory.\footnote{Recently, the 
6879: black hole - string transition has been studied in the context of 
6880: $AdS_5/CFT_4$ correspondence in \cite{Alvarez-Gaume:2006jg}.}
6881: 
6882: 
6883: 
6884: Philosophically, the concept of the phase in the quantum gravity is rather 
6885: obscure. We know that in smaller dimensional field theories or in finite size 
6886: theories, there is no notion of the phase transition. What happens, then, if 
6887: the dimensionality or size of the universe fluctuate as is supposed to be 
6888: the case with the quantum gravity?\footnote{A good example is the de-Sitter 
6889: space, where the quantum gravity is supposed to consist of finite degrees of 
6890: freedom.} Our study only touches a possible hidden nature of the phase 
6891: transition as  non-analytic behaviors of the physical quantities (not 
6892: amplitudes themselves). It would be worth studying further the potential 
6893: origins of the non-analyticity in physical quantities in more general 
6894: situations.
6895: 
6896: 
6897: So far, we have restricted ourselves to $g_s \to 0$ limit throughout this thesis, but finite $g_s$ effects cannot be neglected in any realistic string theory. It is natural to assume that the finite $g_s$ effect sets a cut-off for the emitted closed string energy because it cannot emit energy greater than the tension of the decaying D-brane $\sim 1/g_s$. Therefore, the emission rate roughly behaves as
6898: \begin{eqnarray}
6899: \mathcal{N} \sim \int^{1/g_s} \dd dM N(M) \sqrt{\rho^{(c)}(M)} \ 
6900: \end{eqnarray}
6901: with an explicit cut-off at $1/g_s$. This also means that the radial momentum $p/\sqrt{k}$ should be less than $1/g_s$. Does this constraint smoothen out the phase transition? The saddle point approximation is not accurate as $M$ becomes smaller, so we expect that the phase transition becomes smooth as we introduces $g_s$ corrections. This is consistent with the statement that the ``black hole - string transition" is actually a ``black hole - string crossover". 
6902: 
6903: We have constructed several boundary states for the rolling D-branes in the two-dimensional black hole system. The failure of the uniqueness is physically relevant because in the time-dependent problems in string theory, the boundary conditions should be always imposed in accordance with the physics we are interested in. Mathematically, the different choices of the contour integration give rise to different physics. The tachyon - radion correspondence beautifully connects different solutions (contours) of the rolling tachyon with those for the rolling radion.
6904: 
6905: For the absorbed D-brane boundary condition studied in section \ref{sec:7-2-2}, the dominant infalling closed string radiation (at the Hagedorn temperature) accompanies the outgoing closed string radiation (at the Hawking temperature). The existence of the anomalously small outgoing radiation originates from the boundary condition imposed at the horizon. This reminds us of the closed string Hawking radiation discussed in section \ref{sec:2-4}. Combining the discussion given in section \ref{sec:8-1-4}, we can see that the origin of the thermal-like behaviour of the rolling D-brane radiation is closely related to the boundary condition imposed on the wavefunction (Ishibashi states). It would be interesting to make more precise the relation between the boundary conditions imposed on the string theory and the apparent anomaly as Hawking-like radiation.
6906: 
6907: 
6908: 
6909: Finally, we have discussed the consistency between the unitarity and the 
6910: open-closed duality in the radiative process for the decay of unstable 
6911: D-brane and rolling of accelerated D-brane dynamics in section \ref{sec:5-2} and \ref{sec:8-2} respectively. From ``ab initio" 
6912: derivation in the open string channel, both in Euclidean and Lorentzian 
6913: world-sheet approaches, we have found heretofore overlooked contribution to 
6914: the spectral amplitudes and observables. The contribution is fortuitously 
6915: absent for decay of unstable D-brane, but is present for rolling of 
6916: accelerated D-brane. We have shown that the contribution is imperative for 
6917: ensuring unitarity and optical theorem.
6918: 
6919: Our notion of the unitarity is rather specific, so we have not discussed 
6920: more fundamental questions e.g. about the unitarity of the quantum black hole 
6921: systems associated with Hawking evaporation. The information paradox of the 
6922: black hole system should be resolved within the contex of the string theory 
6923: if it is really a fundamental theory of everything.
6924: 
6925: These three questions raised in this thesis are basic but profound ones that 
6926: people might first come up with when they would like to discuss the 
6927: fundamental properties of the quantum gravity. We have attacked them from 
6928: the concrete examples of the exactly solvable string black hole background. 
6929: At this moment, we do not have complete answers to these questions in the 
6930: tremendously huge string landscape, but we believe that our little step in 
6931: the small corner will ultimately lead to their final answers. 
6932: 
6933: \section*{Acknowledgements}
6934: The author would like to thank all my friends, my family, and my teachers for supporting him to write up this thesis. In particular the author would like to express his sincere thanks to his supervisor Tohru Eguchi for continuous encouragement and advice. Also he would like to express his special thanks to Soo-Jong Rey and Yuji Sugawara for the fruitful collaborations. The most of the main results in this thesis is based on the collaborations with them. He also acknowledges Sylvain Ribault and Yuji Tachikawa for stimulating discussions on the type 2 boundary states and a-theorem violation for generalized conifolds. This research is supported in part by JSPS Research Fellowships for Young Scientists.
6935:  
6936: 
6937: \newpage
6938: 
6939: \appendix\sectiono{Appendices I: Conventions and Useful Formulae}\label{sec:A}
6940: \subsection{conventions}
6941: 
6942: {\bf world-sheet}
6943: 
6944: 
6945: In this thesis, we use $(-,+,+,\cdots,+)$ conventions for target-space metric signature. For the world-sheet coordinate with Lorentzian signature, we use $ -\infty < \tau < \infty$ and $ 0 \le \sigma \le 2\pi$ $(\sigma+2\pi \simeq \sigma)$. The light cone coordinate is defined as
6946: \begin{align}
6947: \begin{cases}
6948: \sigma_+ = \sigma + \tau \cr
6949: \sigma_- = \sigma - \tau 
6950: \end{cases} \ .
6951: \end{align}
6952: We abbreviate their derivatives as $\partial_+ = \frac{\partial}{\partial\sigma_+}$ and $\partial_- = \frac{\partial}{\partial\sigma_-}$.
6953: 
6954: On the other hand, the complex coordinate on the complex plane is defined as
6955: \begin{align}
6956: \begin{cases}
6957: z = x_1 + ix_2 \cr
6958: \bar{z} = x_1 - ix_2 \ 
6959: \end{cases}
6960: \end{align}
6961: We abbreviate their derivatives as $\partial = \frac{\partial}{\partial z}$ and $\bar{\partial} = \frac{\partial}{\partial \bar{z}}$. The integration measure is given by $\dd z^2 \equiv \dd x_1 \dd x_2$.
6962: 
6963: Throughout this thesis, we use the convention $\alpha' = l_s^2 = 2$ as long as otherwise stated. However, in several places, we explicitly write down $\alpha'$ for reader's convenience to compare our results with those in literatures, where different conventions are sometimes used.
6964: 
6965: 
6966: {\bf theta functions with characteristic}
6967: 
6968: \begin{align}
6969: \theta_0(\tau,v) &=\theta_4(\tau,v) = \prod_{m=1}^\infty (1-q^m)(1-zq^{m-1/2})(1-z^{-1}q^{m-1/2}) \cr
6970: \theta_1(\tau,v) &= -2q^{1/4}\sin\pi v \prod_{m=1}^\infty (1-q^m)(1-zq^m)(1-z^{-1}q^m) \cr
6971: \theta_2(\tau,v) &= 2q^{1/4}\cos\pi v \prod_{m=1}^\infty (1-q^m)(1+zq^m)(1+z^{-1}q^m) \cr
6972: \theta_3(\tau,v) &= \prod_{m=1}^\infty (1-q^m)(1+zq^{m-1/2})(1+z^{-1}q^{m-1/2})  \cr
6973: \eta(\tau) &= q^{1/24} \prod_{m=1}^{\infty}(1-q^m) \ ,
6974: \end{align}
6975: where $q = \exp(2\pi i \tau)$ and $z= \exp(2\pi i v)$. Their $S$-modular transformations are
6976: \begin{align}
6977: \theta_0(-1/\tau,v/\tau) &= (-i\tau)^{1/2} \exp(\pi i v^2/\tau) \theta_2(\tau,v) \cr
6978: \theta_1(-1/\tau,v/\tau) &= -(-i\tau)^{1/2} \exp(\pi i v^2/\tau) \theta_1(\tau,v) \cr
6979: \theta_2(-1/\tau,v/\tau) &= (-i\tau)^{1/2} \exp(\pi i v^2/\tau) \theta_0(\tau,v) \cr
6980: \theta_3(-1/\tau,v/\tau) &= (-i\tau)^{1/2} \exp(\pi i v^2/\tau) \theta_3(\tau,v) \cr
6981: \eta(-1/\tau) &=(-i\tau)^{1/2} \eta(\tau) \ .
6982: \end{align}
6983: %where $\tilde{q} = \exp(-2\pi i/\tau)$ and $\tilde{z} = \exp(2\pi i v/\tau)$.
6984: 
6985: The theta functions satisfy the following Riemann quartic identity:
6986: \begin{align}
6987: 2\prod_{a=1}^4 \theta_1(\tau,v'_a) = \prod_{a=1}^4 \theta_3(\tau,v_a) - \prod_{a=1}^4 \theta_2(\tau,v_a) - \prod_{a=1}^4 \theta_0(\tau,v_a) + \prod_{a=1}^4 \theta_1(\tau,v_a) \ , \label{rqi}
6988: \end{align}
6989: where 
6990: \begin{align}
6991: 2v'_1 &= v_1 + v_2 + v_3 + v_4 \ ,& 2v_2'   &= v_1 + v_2 -v_3 -v_4  \ ,\cr
6992: 2v'_3 &= v_1 - v_2 + v_3 -v_4 \ , & 2v_4'  &= v_1 - v_2 - v_ 3 + v_4 \ .
6993: \end{align}
6994: As a corollary, we obtain the abstruse identity of Jacobi:
6995: \begin{eqnarray}
6996: \theta_0^4(\tau,v) + \theta_2^2(\tau,v) = \theta_1^2(\tau,v) + \theta_3^4(\tau,v) \ .
6997: \end{eqnarray}
6998: 
6999: 
7000: {\bf Hypergeometric functions}
7001: 
7002: Gauss's hypergeometric function is defined as
7003: \begin{eqnarray}
7004: F(a_1,a_2;b;z) = \  _2F_1(a_1,a_2;b;z) = \sum_{l=0} \frac{(a_1)_l (a_2)_l}{(b)_l}\frac{x^l}{l!} \ ,
7005: \end{eqnarray}
7006: where 
7007: \begin{eqnarray}
7008: (a)_l \equiv \frac{\Gamma(a+l)}{\Gamma(a)} \ .
7009: \end{eqnarray}
7010: 
7011: The analytic continuation of the hypergeometric function is defined by
7012: \begin{align}
7013: F(\alpha,\beta;\gamma;z) &=
7014: \frac{\Gamma(\gamma)\Gamma(\beta-\alpha)}
7015: {\Gamma(\beta)\Gamma(\gamma-\alpha)}
7016: (-z)^{-\alpha}F(\alpha,\alpha+1-\gamma;
7017: \alpha+1-\beta;1/z) \cr
7018: &+ \frac{\Gamma(\gamma)\Gamma(\alpha-\beta)}
7019: {\Gamma(\alpha)\Gamma(\gamma-\beta)}(-z)^{-\beta}
7020: F(\beta,\beta+1-\gamma;\beta+1-\alpha;1/z)
7021: \label{eq:inv} \ ,
7022: \end{align}
7023: \begin{align}
7024: F(\alpha,\beta;\gamma;z) &=
7025: \frac{\Gamma(\gamma)\Gamma(\gamma-\beta-\alpha)}
7026: {\Gamma(\gamma-\beta)\Gamma(\gamma-\alpha)}
7027: F(\alpha,\beta;
7028: \alpha+\beta+1-\gamma;1-z) \cr
7029: &+ \frac{\Gamma(\gamma)\Gamma(\alpha+\beta-\gamma)}
7030: {\Gamma(\alpha)\Gamma(\beta)}(1-z)^{\gamma-\alpha-\beta}F(\gamma-\alpha,\gamma-\beta;1+\gamma-\alpha-\beta;1-z)
7031: \label{eq:inv2} \ .
7032: \end{align}
7033: If $\gamma = \alpha + \beta + m $ with a certain integer $m$, the above formula should be modified due to the logarithmic singularity at $z=1$. In a particular case ($m=0$), which is interesting to us, the modified formula \cite{transcendental} reads 
7034: \begin{align}
7035: F(\alpha,\beta;\alpha+\beta;z) = \frac{\Gamma(\alpha+\beta)}{\Gamma(\alpha)\Gamma(\beta)} \sum_{n=0}^\infty \frac{a_{(n)}b_{(n)}}{(n!)^2} [h''_n - \log(1-z)] (1-z)^n \ ,
7036: \end{align}
7037: where $h''_n = 2\psi(n+1)-\psi(\alpha+n) - \psi(\beta +n)$ with $\psi(z) \equiv \frac{\Gamma'(z)}{\Gamma(z)}$. 
7038: 
7039: We also introduce the generalized hypergeometric function
7040: \begin{eqnarray}
7041: _3F_2(a_1,a_2,a_3;b_1,b_2;z) = \sum_{l=0} \frac{(a_1)_l (a_2)_l(a_3)_l}{(b_1)_l(b_2)_l}\frac{x^l}{l!} \ ,
7042: \end{eqnarray}
7043: whose asymptotic expansion (as $z\to \infty$) is given by
7044: \begin{align}
7045: &_3F_2(a_1,a_2,a_3;b_1,b_2;z) \cr &
7046: = \frac{\Gamma(b_1)\Gamma(b_2)}{\Gamma(a_1)\Gamma(a_2)\Gamma(a_3)} \sum_{k=1}^3 \frac{\Gamma(a_k)\prod_{j=1;j\neq k}^3(a_j-a_k)}{\prod_{j=1}^2 \Gamma(b_j-a_k)}(-z)^{-a_k}\left[1 + O(z^{-1})\right] \ .
7047: \end{align}
7048: 
7049: \subsection{$SL(2;\br)$ current algebra}\label{a-1}
7050: We will collect useful facts about the $SL(2;\br)$ current algebra and fix our notations. We begin with the zero mode. Our conventions for the commutation relations  of the $SL(2;\br)$ algebra are
7051: \begin{align}
7052: [J^1,J^2] = -i J^3 \ , [J^2,J^3] = i J^1 \ , [J^3,J^1] = iJ^2 \ .
7053: \end{align}
7054: We will also introduce
7055: \begin{align}
7056: J^{\pm} = J^1 \pm iJ^2 \ ,
7057: \end{align}
7058: which gives 
7059: \begin{align}
7060: [J^3, J^{\pm}] = \pm J^{\pm} \ , [J^+,J^{-}] = -2J^3 \ . 
7061: \end{align}
7062: The quadratic Casimir is defined as
7063: \begin{align}
7064: C_2 = (J^1)^2 + (J^2)^2 - (J^3)^2 \equiv -j(j-1) \ .
7065: \end{align}
7066: We summarize the unitary (irreducible) representations of the $SL(2;\br)$ algebra parametrized by $j$
7067: 
7068: \begin{enumerate}
7069: 	\item Principal discrete representations (highest or lowest weight states): they contain the state that is annihilated by $J^+$ or $J^-$ respectively. 
7070: Their modules are generated as
7071: \begin{align}
7072: D^+ = \left\{ |j;m\rangle : m = -j , -j+1, -j+2, \cdots\right\} \ ,
7073: \end{align}
7074: and
7075: \begin{align}
7076: D^- = \left\{ |j;m\rangle : m = j , j-1, j-2, \cdots\right\} \ ,
7077: \end{align}
7078: obtained by acting $J^-$ or $J^+$ on the highest or lowest weight state.
7079: The representation is unitary when $j \le -\frac{1}{2}$.
7080: 	\item Principal continuous series: they are realized by
7081: \begin{align}
7082: C_j^{\alpha} = \left\{|j,\alpha;m\rangle : m = \alpha, \alpha \pm1, \alpha \pm 2, \cdots \right\}
7083: \end{align}
7084: where $J^3|j,\alpha;m\rangle = m |j,\alpha;m\rangle$. Without loss of generality, we take $0\le \alpha <1$. The representation is unitary when $j=-1/2 + is $ with $s\in \br^+$.
7085: 
7086: 	\item Complementary representations: similar to the continuous series but with real $j$. They are unitary when $-1 < j <-1/2$ and $-j-1/2<|\alpha-1/2|$.
7087: 	\item Identity representation: this is the trivial representation with $j=-1$.
7088: \end{enumerate}
7089: 
7090: 
7091: 
7092: Let us move to the affine current algebra $\widehat{SL(2;\br)}$. The relevant OPEs are
7093: \begin{align}
7094: \begin{cases} j^3(z)j^3(0) \sim -\frac{\kappa}{2z^2} \cr
7095: j^3(z) j^{\pm}(0) \sim \pm \frac{1}{z}j^{\pm}(0) \cr
7096: j^+(z)j^{-}(0) \sim \frac{\kappa}{z^2} - \frac{2}{z}j^{3}(0) 
7097: \end{cases} \ .
7098: \end{align}
7099: Corresponding affine Lie algebra is given by
7100: \begin{align}
7101: [J^3_n,J^3_m] &= -\frac{k}{2}n \delta_{m,-n} \ , \cr
7102: [J^3_n,J^{\pm}_m] & = \pm J^\pm_{n+m} \ , \cr
7103: [J^+_n,J^{-}_m] & = -2J^3_{n+m} + kn\delta_{m,-n} \ .
7104: \end{align}
7105: The energy momentum tensor is given by the Sugawara form:
7106: \begin{align}
7107: T(z) = \frac{1}{\kappa-2}(j^1j^1 + j^2 j^2 -j^3j^3)(z) \ 
7108: \end{align}
7109: with the central charge $c=3 + \frac{6}{\kappa-2}$.
7110: 
7111: We summarize the characters of the unitary  representations of $\widehat{SL(2;\br)}$ current algebra. The character is defined as $\mathrm{Tr} q^{L_0} y^{J_0^3} $ with $q\equiv e^{2\pi i\tau}$, $y\equiv e^{2\pi i u}$.
7112: \begin{enumerate}
7113: 	\item Principal discrete representations (highest or lowest weight states): the characters can be written as
7114: \begin{align}
7115: \chi_j^{\pm}(\tau,u) = \pm i \frac{q^{-\frac{1}{\kappa-2}(j-\frac{1}{2})^2}y^{\pm (j-\frac{1}{2})}}{\theta_1(\tau,u)} \ .
7116: \end{align}
7117: 
7118: 	\item Principal continuous series: the character is given by
7119: \begin{align}
7120: \chi_{s,\alpha}(\tau,u) = \frac{q^{\frac{s^2}{\kappa-2}}}{\eta(\tau)^3} \sum_{n\in \bz} y^{n+\alpha} \ .
7121: \end{align}
7122: 
7123: 	\item Complementary representations: the character is given by
7124: \begin{align}
7125: \chi_{j,\alpha}(\tau,u) = \frac{q^{\frac{(j-\frac{1}{2})^2}{\kappa-2}}}{\eta(\tau)^3} \sum_{n\in \bz} y^{n+\alpha} \ .
7126: \end{align}
7127: 
7128: 	\item Identity representation: the character is given by
7129: \begin{align}
7130: \chi_0(\tau,u) = i \frac{q^{-\frac{1}{4(\kappa-2)}}y^{-1/2}(1-y)}{\theta_1(\tau,u)} \ .
7131: \end{align}
7132: It is possible to construct more general representations of the current algebra by using ($n$-units of) the spectral flows
7133: \begin{align}
7134: j_m^3 \to j_m^3 - \frac{\kappa}{2}n \delta_{m,0} \ ,  \ \ j^{\pm}_m \to j^{\pm}_{m\pm n} \ , \ \ L_m \to L_m + nj^3_m - \frac{\kappa}{4}n^2 \delta_{m,0} \ . 
7135: \end{align}
7136: Unlike in the case of the compact group, we actually obtain new representations, but their conformal dimensions are typically unbounded below.
7137: \end{enumerate}
7138: 
7139: \subsection{Coordinate on $SL(2;\br)$}\label{a-1-2}
7140: The Euler angle parametrization $g(r,t,\phi) \in SL(2;\br)$ suitable for the Euclidean coset is given by
7141: \begin{eqnarray}
7142: e^{i\sigma_1\frac{t-\phi}{2}} e^{r\sigma_2}e^{i\sigma_1\frac{t+\phi}{2}} 
7143: = \begin{pmatrix} \cos t \cosh r - \sin\phi \sinh r & \cosh r \sin t + \cos \phi \sinh r \cr
7144: -\cosh r \sin t + \cos \phi \sinh r& \cos t \cosh r + \sin\phi \sinh r 
7145: \end{pmatrix} \ ,
7146: \end{eqnarray}
7147: where $0<t\le 2\pi$, $ 0\le r$ and $0<\phi \le 2\pi$ for the single cover of $SL(2;\br)$. 
7148: 
7149: On the other hand, for the Lorentzian coset, a useful Euler angle parametrization $g(r,t_L,t_R)$ is given by 
7150: \begin{eqnarray}
7151:  e^{\sigma_3 t_L} e^{r\sigma_2}e^{\sigma_3 t_R} = \begin{pmatrix} e^{t_L+t_R} \cosh r & e^{t_L-t_R} \sinh r \cr
7152: e^{-t_L+t_R} \sinh r & e^{-t_L-t_R} \cosh r \end{pmatrix} \ , 
7153: \end{eqnarray}
7154: where $t_L$ and $t_R$ are noncompact.\footnote{The parametrization does {\it not} cover the whole $SL(2;\br)$ manifold.}
7155: 
7156: To connect it with the $AdS_3$ space, it is customary to introduce the parametrization
7157: \begin{eqnarray}
7158: g = \begin{pmatrix} x^0 + x^2 & x^1 + x^3 \cr x^1-x^3 & x^0-x^2 \end{pmatrix} \ 
7159: \end{eqnarray}
7160:  so that we can see the $SL(2;\br)$ group as a hyperbola 
7161: \begin{eqnarray}
7162: (x^0)^2 - (x^1)^2 - (x^2)^2 + (x^3)^2 = 0 \ 
7163: \end{eqnarray}
7164: embedded in Minkowski space $(x^0,x^1,x^2,x^3)$ with signature $(-,+,+,-)$.
7165: 
7166: \subsection{Frequently used formulae} 
7167: 
7168: \begin{eqnarray}
7169: \Gamma(z+1) = z\Gamma(z) \ .
7170: \end{eqnarray}
7171: \begin{eqnarray}
7172: \Gamma(z)\Gamma(1-z) = \frac{\pi}{\sin\pi z} \ .
7173: \end{eqnarray}
7174: \begin{eqnarray}
7175: \Gamma(\frac{1}{2}+z)\Gamma(\frac{1}{2}-z) = \frac{\pi}{\cos\pi z} \ .
7176: \end{eqnarray}
7177: \begin{eqnarray}
7178: \Gamma(2z) = (2\pi)^{-1/2} 2^{2z-1/2} \Gamma(z) \Gamma(z+1/2) \ .
7179: \end{eqnarray}
7180: \begin{align}
7181: \int_0^\infty dx \frac{x^c}{\sqrt{x^2 + a^2}} = \frac{a^c \Gamma(-\frac{c}{2})\Gamma(\frac{1+c}{2})}{2\sqrt{\pi}} \ .
7182: \end{align}
7183: \begin{align}
7184: & \hspace{-5mm} \int_{-\frac{\pi}{2}}^{\frac{\pi}{2}} (2 \cos
7185: \theta)^{a-1} e^{i b \theta} \dd \theta = \pi \frac{\Gamma(a)}
7186: {\Gamma\left(\frac{1}{2}+\frac{a+b}{2}\right)
7187: \Gamma\left(\frac{1}{2}+\frac{a-b}{2}\right)}~,~~~ (\Re\,
7188: a>0~,~~\left|\Re\, b \right|< \Re\, a + 1)~,
7189: \label{formula 1} \\
7190: & \hspace{-5mm} \int_{-\infty}^{\infty} (2 \cosh t)^{a-1} e^{i b t}
7191: \dd t = \frac{1}{2} B\left(\frac{1}{2}-\frac{a+ib}{2},
7192: \frac{1}{2}-\frac{a-ib}{2}\right) \equiv \frac{1}{2} \frac
7193: {\Gamma\left(\frac{1}{2}-\frac{a+ib}{2}\right)
7194: \Gamma\left(\frac{1}{2}-\frac{a-ib}{2}\right)} {\Gamma(1-a)}~,\nn &
7195: \hspace{10cm} (\Re\, a <1~,~~ \left|\Im \, b \right| < 1- \Re\, a)~.
7196: \label{formula 2}
7197: \end{align}
7198: The integral \eqref{formula 2} follows from the more general formula:
7199: \begin{align}
7200: \int_0^{\infty} \frac{\cosh(2at)}{\cosh^{2\beta} (pt)} \dd t =
7201: 4^{\beta-1} p^{-1} B\left(\beta+\frac{a}{p},
7202: \beta-\frac{a}{p}\right)~,~~~ (p>0,
7203: ~~\Re\,\left(\beta\pm\frac{a}{p}\right)>0)~,
7204: \end{align}
7205: given in \cite{transcendental}.
7206: 
7207: 
7208: 
7209: 
7210: 
7211: 
7212: 
7213: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
7214: 
7215: %\appendixname{~ B }
7216: \subsection{Proof of \eqref{evaluation overlap phi} and \eqref{evaluation overlap phi2}}\label{mini}
7217:  
7218: 
7219: Here we would like to evaluate explicitly the integral
7220: \eqref{evaluation overlap phi} for any $\rho_0$ (strictly speaking, we
7221: need to assume $\sinh \rho_0>1$). We begin with series expansion of
7222: the hypergeometric function in $\phi^p_n(\rho,\theta)$:
7223: \begin{align}
7224: &
7225: F\left(\frac{1}{2}+\frac{ip+n}{2},\frac{1}{2}+\frac{ip-n}{2};
7226: ip+1;
7227: -\frac{\cos^2 \theta}{\sinh^2 \rho_0}\right)\nn
7228: &=\sum_{\ell=0}^\infty \frac{\Gamma(ip+1)}
7229: {\Gamma(\frac{1}{2}+\frac{ip+n}{2})
7230: \Gamma(\frac{1}{2}+\frac{ip-n}{2})}
7231: \frac{\Gamma(\frac{1}{2}+\frac{ip+n}{2}+\ell)
7232: \Gamma(\frac{1}{2}+\frac{ip-n}{2}+\ell)}
7233: {\Gamma(ip+1+\ell)}\frac{(-1)^\ell}{\ell !}
7234: \left(\frac{\cos \theta}{\sinh \rho_0}\right)^{2\ell}.
7235: \label{expansion}
7236: \end{align}
7237: Using the formula (\ref{formula 1}), we can perform, in the
7238: $\ell$-th sector, the integral (\ref{evaluation overlap phi}) as
7239: \begin{align}
7240: \Psi_\ell=g(\ell)\int_{-\frac{\pi}{2}}^{\frac{\pi}{2}} \dd\theta \,
7241: e^{in\theta}(\cos \theta)^{ip-1+ 2\ell} =\frac{g(\ell)}{2^{ip-1
7242: +2\ell}}\cdot \frac{\pi\Gamma(ip-1+2\ell+1)}
7243: {\Gamma(\frac{1}{2}+\ell+\frac{ip+n}{2})
7244: \Gamma(\frac{1}{2}+\ell+\frac{ip-n}{2})}, \nonumber
7245: \end{align}
7246: where $g(\ell)$ refers to
7247: \begin{equation}
7248: g(\ell)=\frac{(-1)^\ell}{\ell !}
7249: (\sinh \rho_0)^{-ip -2\ell}\frac{\Gamma(ip+1)}
7250: {\Gamma(\frac{1}{2}+\frac{ip+n}{2})
7251: \Gamma(\frac{1}{2}+\frac{ip-n}{2})}
7252: \frac{\Gamma(\frac{1}{2}+\frac{ip+n}{2}+\ell)
7253: \Gamma(\frac{1}{2}+\frac{ip-n}{2}+\ell)}{\Gamma(ip+1+\ell)}.
7254: \end{equation}
7255: Then the total integral (\ref{evaluation overlap phi}) is
7256: \begin{equation}
7257: \sum_{\ell=0}^\infty \Psi_\ell
7258: =\frac{\pi}{2^{ip-1}(\sinh \rho_0)^{ip}}\frac{\Gamma(ip+1)}
7259: {\Gamma(\frac{1}{2}+\frac{ip+n}{2})
7260: \Gamma(\frac{1}{2}+\frac{ip-n}{2})}\sum_{\ell=0}^\infty
7261: \frac{(-1)^\ell}{\ell !}\frac{1}{2^{2\ell}(\sinh \rho_0)^{2\ell}}
7262: \frac{\Gamma(ip+2\ell)}{\Gamma(ip+1+\ell)}.
7263: \end{equation}
7264: We can rewrite the summation into a hypergeometric function
7265: by using
7266: %\footnote{Recall the formula
7267: %\begin{equation}
7268: %\Gamma(2z)
7269: %=\frac{2^{2z}}{2\sqrt{\pi}}\Gamma(z)\Gamma(z+\frac{1}{2}).
7270: %\label{calc0}
7271: %\end{equation}}
7272: \begin{equation}
7273: \Gamma(ip+2\ell)=\frac{2^{ip-1+2\ell}}{\sqrt{\pi}}
7274: \Gamma\left(\frac{ip}{2}+\ell\right)
7275: \Gamma\left(\frac{1}{2}+\frac{ip}{2}+\ell\right)~.
7276: \end{equation}
7277: and then obtain
7278: \begin{equation}
7279: \sum_{\ell=0}^\infty \Psi_\ell=\frac{\sqrt{\pi}}
7280: {(\sinh \rho_0)^{ip}}\frac{\Gamma(\frac{ip}{2})
7281: \Gamma(\frac{1}{2}+\frac{ip}{2})}
7282: {\Gamma(\frac{1}{2}+\frac{ip+n}{2})
7283: \Gamma(\frac{1}{2}+\frac{ip-n}{2})}
7284: F\left(\f{ip}{2},\frac{1}{2}+\frac{ip}{2}; ip+1 ;
7285: -\frac{1}{\sinh^2 \rho_0}\right).
7286: \label{calc1}
7287: \end{equation}
7288: Making use of the formula
7289: %(e.g. see p. 59 in the Vol. 3
7290: %of {\it Iwanami's book})
7291: \begin{align}
7292: &F\left(2\alpha,2\beta;\alpha+\beta+\frac{1}{2};z\right)=
7293: F\left(\alpha,\beta;\alpha+\beta+\frac{1}{2};4z(1-z)\right).
7294: \label{calc2} \\
7295: & \hskip5cm |z|<\frac{1}{2},\quad |z(1-z)|<\frac{1}{4}~ \nonumber
7296: \end{align}
7297: %This formula is valid for
7298: %\begin{equation}
7299: %|z|<\frac{1}{2},\quad |z(1-z)|<\frac{1}{4}.
7300: %\end{equation}
7301: %The hypergeometric function in Eq.(\ref{calc1}) has the form of
7302: %the R.H.S in Eq.(\ref{calc2}).
7303: %Taking care of the region of $z$, we obtain
7304: %\begin{equation}
7305: %4z(1-z)=-\f{1}{\sinh^2 \rho_0}
7306: %\quad \rightarrow \quad z=\frac{1}{2}-
7307: %\frac{\cosh \rho_0}{2\sinh \rho_0}.
7308: %\end{equation}
7309: we find that
7310: \begin{equation}
7311: \sum_{\ell=0}^\infty \Psi_\ell=\frac{\sqrt{\pi}}
7312: {(\sinh \rho_0)^{ip}}\frac{\Gamma(\frac{ip}{2})
7313: \Gamma(\frac{1}{2}+\frac{ip}{2})}
7314: {\Gamma(\frac{1}{2}+\frac{ip+n}{2})
7315: \Gamma(\frac{1}{2}+\frac{ip-n}{2})}
7316: F\left(ip,ip+1,ip+1;\frac{1}{2}-
7317: \frac{\cosh \rho_0}{2\sinh \rho_0}\right).
7318: \end{equation}
7319: Note that the second and third arguments
7320: of the hypergeometric function are the same.
7321: The function is thus simplified as
7322: \begin{equation}
7323: F\left(ip,ip+1,ip+1;\frac{1}{2}-
7324: \frac{\cosh \rho_0}{2\sinh \rho_0}\right)=
7325: \left(\frac{\sinh \rho_0 +\cosh \rho_0}
7326: {2\sinh \rho_0}\right)^{-ip}
7327: =\left(2e^{-\rho_0}\sinh \rho_0\right)^{ip}~,
7328: \end{equation}
7329: because of the relation
7330: \begin{equation}
7331: (1-z)^{\alpha}=F(-\alpha,\beta;\beta;z)~.
7332: \end{equation}
7333: In this way, we finally obtain
7334: \begin{align}
7335: & \sum_{\ell=0}^\infty \Psi_\ell=\sqrt{\pi}
7336: e^{-ip \rho_0 }2^{ip}
7337: \frac{\Gamma(\frac{ip}{2})\Gamma(\frac{1}{2}+\frac{ip}{2})}
7338: {\Gamma(\frac{1}{2}+\frac{ip+n}{2})\Gamma(\frac{1}{2}
7339: +\frac{ip-n}{2})}
7340: = \frac{2\pi \,\Gamma(ip)}
7341: {\Gamma(\frac{1}{2}+\frac{ip+n}{2})
7342: \Gamma(\frac{1}{2}+\frac{ip-n}{2})}e^{-ip\rho_0}~,
7343: \end{align}
7344: %Finally after using the formula (\ref{calc0}) again, we obtain
7345: %\begin{equation}
7346: %\sum_{\ell=0}^\infty \Psi_\ell=\frac{2\pi \,\Gamma(ip)}
7347: %{\Gamma(\frac{1}{2}+\frac{ip+n}{2})
7348: %\Gamma(\frac{1}{2}+\frac{ip-n}{2})}e^{-ip\rho_0}.
7349: %\end{equation}
7350: and this is the desired formula.
7351: %\section*{Appendix B}\label{mini2}
7352: 
7353: %\setcounter{equation}{0}
7354: %\def\theequation{B.\arabic{equation}}
7355: %\indent
7356: 
7357: In a quite similar manner, we can prove
7358: \eqref{evaluation overlap phi2}. We begin with series expansion of
7359: the hypergeometric function in $\phi_{pw}(\rho,\theta)$:
7360: \begin{align}
7361: &
7362: F\left(\frac{1}{2}+\frac{ip+kw}{2},\frac{1}{2}+\frac{ip-kw}{2};
7363: ip+1;
7364: \frac{\cos^2 \theta}{\cosh^2 r_0}\right)\nn
7365: &=\sum_{\ell=0}^\infty \frac{\Gamma(ip+1)}
7366: {\Gamma(\frac{1}{2}+\frac{ip+kw}{2})
7367: \Gamma(\frac{1}{2}+\frac{ip-kw}{2})}
7368: \frac{\Gamma(\frac{1}{2}+\frac{ip+kw}{2}+\ell)
7369: \Gamma(\frac{1}{2}+\frac{ip-kw}{2}+\ell)}
7370: {\Gamma(ip+1+\ell)}\frac{1}{\ell !}
7371: \left(\frac{\cos \theta}{\cosh r_0}\right)^{2\ell}.
7372: \label{expansion}
7373: \end{align}
7374: Using the formula (\ref{formula 1}), we can perform, in the
7375: $\ell$-th sector, the integral (\ref{evaluation overlap phi}) as
7376: \begin{align}
7377: \Psi_\ell=g(\ell)\int_{-\frac{\pi}{2}}^{\frac{\pi}{2}} \dd\theta \,
7378: e^{ikw\theta}(\cos \theta)^{ip-1+ 2\ell} =\frac{g(\ell)}{2^{ip-1
7379: +2\ell}}\cdot \frac{\pi\Gamma(ip-1+2\ell+1)}
7380: {\Gamma(\frac{1}{2}+\ell+\frac{ip+kw}{2})
7381: \Gamma(\frac{1}{2}+\ell+\frac{ip-kw}{2})}, \nonumber
7382: \end{align}
7383: where $g(\ell)$ refers to
7384: \begin{equation}
7385: g(\ell)=\frac{1}{\ell !}
7386: (\cosh r_0)^{-ip -2\ell}\frac{\Gamma(ip+1)}
7387: {\Gamma(\frac{1}{2}+\frac{ip+kw}{2})
7388: \Gamma(\frac{1}{2}+\frac{ip-kw}{2})}
7389: \frac{\Gamma(\frac{1}{2}+\frac{ip+kw}{2}+\ell)
7390: \Gamma(\frac{1}{2}+\frac{ip-kw}{2}+\ell)}{\Gamma(ip+1+\ell)}.
7391: \end{equation}
7392: Then the total integral (\ref{evaluation overlap phi}) is
7393: \begin{equation}
7394: \sum_{\ell=0}^\infty \Psi_\ell
7395: =\frac{\pi}{2^{ip-1}(\cosh r_0)^{ip}}\frac{\Gamma(ip+1)}
7396: {\Gamma(\frac{1}{2}+\frac{ip+kw}{2})
7397: \Gamma(\frac{1}{2}+\frac{ip-kw}{2})}\sum_{\ell=0}^\infty
7398: \frac{1}{\ell !}\frac{1}{2^{2\ell}(\cosh r_0)^{2\ell}}
7399: \frac{\Gamma(ip+2\ell)}{\Gamma(ip+1+\ell)}.
7400: \end{equation}
7401: We can rewrite the summation into a hypergeometric function
7402: by using
7403: \begin{equation}
7404: \Gamma(ip+2\ell)=\frac{2^{ip-1+2\ell}}{\sqrt{\pi}}
7405: \Gamma\left(\frac{ip}{2}+\ell\right)
7406: \Gamma\left(\frac{1}{2}+\frac{ip}{2}+\ell\right)~.
7407: \end{equation}
7408: and then obtain
7409: \begin{equation}
7410: \sum_{\ell=0}^\infty \Psi_\ell=\frac{\sqrt{\pi}}
7411: {(\cosh r_0)^{ip}}\frac{\Gamma(\frac{ip}{2})
7412: \Gamma(\frac{1}{2}+\frac{ip}{2})}
7413: {\Gamma(\frac{1}{2}+\frac{ip+kw}{2})
7414: \Gamma(\frac{1}{2}+\frac{ip-kw}{2})}
7415: F\left(\f{ip}{2},\frac{1}{2}+\frac{ip}{2}; ip+1 ;
7416: \frac{1}{\cosh^2 r_0}\right).
7417: \label{calc1}
7418: \end{equation}
7419: Making use of the formula
7420: \begin{align}
7421: &&F\left(2\alpha,2\beta;\alpha+\beta+\frac{1}{2};z\right)=
7422: F\left(\alpha,\beta;\alpha+\beta+\frac{1}{2};4z(1-z)\right).
7423: \label{calc2} \\
7424: && \hskip5cm |z|<\frac{1}{2},\quad |z(1-z)|<\frac{1}{4}~ \nonumber
7425: \end{align}
7426: we find that
7427: \begin{equation}
7428: \sum_{\ell=0}^\infty \Psi_\ell=\frac{\sqrt{\pi}}
7429: {(\sinh r_0)^{ip}}\frac{\Gamma(\frac{ip}{2})
7430: \Gamma(\frac{1}{2}+\frac{ip}{2})}
7431: {\Gamma(\frac{1}{2}+\frac{ip+n}{2})
7432: \Gamma(\frac{1}{2}+\frac{ip-n}{2})}
7433: F\left(ip,ip+1,ip+1;\frac{1}{2}(1-\tanh r_0)\right).
7434: \end{equation}
7435: Note that the second and third arguments
7436: of the hypergeometric function are the same.
7437: The function is thus simplified as
7438: \begin{equation}
7439: F\left(ip,ip+1,ip+1;\frac{1}{2}(1-\tanh r_0)\right)
7440: =\left(2e^{-r_0}\cosh r_0\right)^{ip}~,
7441: \end{equation}
7442: because of the relation
7443: \begin{equation}
7444: (1-z)^{\alpha}=F(-\alpha,\beta;\beta;z)~.
7445: \end{equation}
7446: In this way, we finally obtain
7447: \begin{align}
7448: && \sum_{\ell=0}^\infty \Psi_\ell=\sqrt{\pi}
7449: e^{-ip r_0 }2^{ip}
7450: \frac{\Gamma(\frac{ip}{2})\Gamma(\frac{1}{2}+\frac{ip}{2})}
7451: {\Gamma(\frac{1}{2}+\frac{ip+kw}{2})\Gamma(\frac{1}{2}
7452: +\frac{ip-kw}{2})}
7453: = \frac{2\pi \,\Gamma(ip)}
7454: {\Gamma(\frac{1}{2}+\frac{ip+kw}{2})
7455: \Gamma(\frac{1}{2}+\frac{ip-kw}{2})}e^{-ipr_0}~,
7456: \end{align}
7457: and this is the desired formula.
7458: 
7459: 
7460: \section{Appendix II: Miscellaneous Topics}\label{sec:B}
7461: 
7462: \subsection{Partition functions}\label{part}
7463: The modular invariant torus partition function is of critical importance in closed string theory to read the spectrum of a given background.  In this appendix, we collect partition functions of several CFTs that are relevant for our discussions.
7464: 
7465: Our main focus is the partition function for the Lorentzian two-dimensional black hole. The partition function for the two-dimensional black hole, however, suffers some subtleties because of the Lorentzian signature of the target space and the divergence coming from the non-compact target space. 
7466: 
7467: With these subtleties  in mind, our starting point is the (twisted) partition function for the $\mathbb{H}_3^+$ model. For a time-being, we restrict ourselves to the bosonic CFT. Since the Euclidean action is bounded, the direct path integral computation is possible \cite{Gawedzki:1988nj,Gawedzki:1991yu}. For the vector gauging
7468: \begin{align}
7469: Z^{\mathbb{H}^3_+}_{(V)}(\tau,u) \equiv \int \dd g e^{-kS^{\mathbb{H}_3^+}_{(V)}(g,h^u,h^{u\dagger})} = \frac{e^{\frac{u_2^2}{\tau_2}}}{\sqrt{\tau_2}|\theta_1(\tau,u)|^2 }\ .
7470: \end{align}
7471: For the axial gauging
7472: \begin{align}
7473: Z^{\mathbb{H}^3_+}_{(A)}(\tau,u) \equiv \int \dd g e^{-kS^{\mathbb{H}_3^+}_{(A)}(g,h^u,h^{u\dagger})} = \frac{e^{\frac{u_2^2}{\tau_2}-\pi k\frac{|u|^2}{\tau_2}}}{\sqrt{\tau_2}|\theta_1(\tau,u)|^2 }\ .
7474: \end{align}
7475: In the $u\to0$ limit, both expressions coincide and (formally) modular invariant.\footnote{Because of the divergence as $u\to 0$, these expressions contain rather poor information to read the spectrum (e.g. the partition function is $k$ independent). Actually, the resurrection of the $k$ dependence is one of the key issues to extract contributions from the discrete states and give the improved unitarity bound. In order to do this, we should actually know the central charge of the model independently since the torus partition function does not know {\it a priori} the shift of the central charge coming from the linear dilaton coupled to the curvature of the world-sheet.} 
7476: 
7477: The partition function for the Euclidean two-dimensional black hole from the axial coset of the $\mathbb{H}^3_+$ model (denoted by ${\mathbb{H}^{3(A)}_+/\mathbb{R}}$) is 
7478: \begin{align}
7479: Z_{\mathbb{H}^{3(A)}_{+}/\br} = \int_{\Sigma} \frac{\dd u^2}{\tau_2} \frac{e^{\frac{u_2^2}{\tau_2}}}{\sqrt{\tau_2}|\theta_1(\tau,u)|^2 } \sqrt{\tau_2}|\eta(\tau)|^2 \sum_{m,\omega\in \bz} e^{-\frac{\pi k}{\tau_2}|\omega\tau - m + u|^2} \ , \label{hpz}
7480: \end{align}
7481: where the integration is over the Jacobian torus $\int_0^1 ds_1 \int_0^1 ds_2$ with $u=s_1\tau-s_2$. The appearance of the (twisted) partition of the compactified boson (with radius $R^2 = k$) 
7482: \begin{align}
7483: Z_{\sqrt{k}}(\tau,u) = \sqrt{\tau_2}|\eta(\tau)|^2 \sum_{m,\omega\in \bz} e^{-\frac{\pi k}{\tau_2}|\omega\tau - m + u|^2}
7484: \end{align}
7485: suggests an asymptotic geometry of the cigar (with the same radius).
7486: Note that the summation over $m,\omega$ can be combined into
7487: \begin{align}
7488: Z_{\mathbb{H}^{3(A)}_{+}/\br} =  \int_{\br^2} \frac{\dd u^2}{\tau_2}\frac{e^{\frac{u_2^2}{\tau_2}-\frac{\pi k}{\tau_2}|u|^2}  }{\sqrt{\tau_2}|\theta_1(\tau,u)|^2 } \sqrt{\tau_2}|\eta(\tau)|^2  \ .
7489: \end{align} 
7490: Due to the fact that ${\mathbb{H}^{3(A)}_{+}/\br}  = SL(2;\br)^{(A)}/U(1)$, one can interpret \eqref{hpz} as the partition function of the $SL(2;\br)^{(A)}/U(1)$ axial coset model.\footnote{In the Euclidean axial coset model, there is no distinction between the single cover of $SL(2;\br)$ and the universal cover of the $SL(2;\br)$.}
7491: 
7492: The partition function for the Lorentzian coset is rather subtle, and we should resort to analytic continuations. Instead of gauging compact subgroup generated by $J^3$, we gauge the noncompact subgroup generated by $J^2$. Effectively, we can Wick rotate $J^3 \to iJ^3$. The axially $J^2$ twisted partition function for (the universal cover of) $SL(2;\br)_{(\infty)}^{(A)}$ could be obtained from the analytic continuation $(u\to iv)$ as
7493: \begin{align}
7494: Z^{SL(2;\br)}_{(A)}(\tau,v) \equiv \int \dd g e^{-kS^{H_3^+}_{(V)}(g,h^v,h^{v\dagger})} = \frac{e^{\frac{v_1^2}{\tau_2}}}{\sqrt{\tau_2}|\theta_1(\tau,v)|^2 } \ .
7495: \end{align}
7496: The partition function for the Lorentzian two-dimensional black hole (after an analytic continuation) could be written as
7497: \begin{align}
7498: Z_{\mathbb{H}^{3(A)}_{+}/\br} =  \int_{\br^2} \frac{\dd v^2}{\tau_2}\frac{e^{\frac{v_1^2}{\tau_2}-\frac{\pi k}{\tau_2}|v|^2}  }{\sqrt{\tau_2}|\theta_1(\tau,iv)|^2 } \sqrt{\tau_2}|\eta(\tau)|^2  \ ,
7499: \end{align}
7500: which agrees with the proposal made in \cite{Israel:2003ry} from a different perspective.
7501: 
7502: 
7503: \subsection{T-duality and Buscher's rule}\label{busc}
7504: We summarize Buscher's rule \cite{Buscher:1987sk,Buscher:1987qj} for T-duality:
7505: \begin{align}
7506: \tilde{G}_{00} &= \frac{1}{G_{00}} \cr
7507: \tilde{G}_{0i} &= \frac{B_{0i}}{G_{00}} \cr
7508: \tilde{B}_{0i} &= \frac{G_{0i}}{G_{00}} \cr
7509: \tilde{G}_{ij} &= G_{ij} - \frac{G_{0i}G_{0j}-B_{0i}B_{0j}}{G_{00}} \cr
7510: \tilde{B}_{ij} &= B_{ij} -\frac{G_{0i}B_{0j}-B_{0i}G_{0i}}{G_{00}} \cr
7511: \tilde{\Phi} &= \Phi - \frac{1}{4}\log\left(\frac{G_{00}}{\tilde{G}_{00}}\right) \ ,
7512: \end{align}
7513: where $0$ is the direction to be T-dualized.
7514: 
7515: \subsection{$\mathcal{N}=2$ superconformal algebra and spectral flow}\label{SCA2}
7516: OPEs for the $\mathcal{N}=2$ SCA are summarized as
7517: \begin{align}
7518: T(z) T(0) &\sim \frac{c}{2z^4} + \frac{2}{z^2}T(0) + \frac{1}{z}\partial T(0) \cr
7519: T(z) G^{\pm}(0) &\sim \frac{3}{2z^2}G^{\pm}(0) + \frac{1}{z}\partial G^{\pm}(0) \cr
7520: T(z) J(0) &\sim \frac{1}{z^2} J(0) + \frac{1}{z} \partial J(0) \cr
7521: G^+(z)G^{-}(0) &\sim \frac{2c}{3z^3}+\frac{2}{z^2}J(0) + \frac{2}{z} T(0) + \frac{1}{z} \partial J(0) \cr
7522: G^{\pm}(z)G^{\pm}(0) &\sim 0 \cr
7523: J(z)G^{\pm}(0) &\sim \pm G^{\pm}(0) \cr
7524: J(z)J(0) &\sim \frac{c}{3z^2} \ .
7525: \end{align}
7526: Correspondingly the mode expansion yields 
7527: \begin{align}
7528: [L_m,L_n] &= (m-n)L_{m+n} + \frac{c}{12}(m^3-m)\delta_{m,-n} \cr
7529: [L_m,G^{\pm}_r] &= \left(\frac{m}{2}-r\right)G^{\pm}_{m+r} \cr
7530: [L_m,J_n] &= -n J_{m+n} \cr
7531: \{G_r^{+},G_{s}^{-}\} &= 2L_{r+s} + (r-s)J_{r+s} + \frac{c}{3}\left(r^2 -\frac{1}{4}\right) \delta_{r,-s} \cr
7532: \{G_r^{\pm},G_s^{\pm}\} &= 0 \cr
7533: [J_n,G^{\pm}_r] &= \pm G_{r+n}^{\pm} \cr
7534: [J_m,J_n] &= \frac{c}{3}m\delta_{m,-n} \ ,
7535: \end{align}
7536: where $r,s$ are integers for the R-sector and half-integers for the NS-sector.
7537: The $\mathcal{N}=2$ superconformal algebra admits an isomorphism known as spectral flow:
7538: \begin{align}
7539:  U_{\eta}L_{m}U_\eta^{-1} &\to L_{m} + \eta J_m +\frac{\eta^2 c}{6}\delta{m,0} \cr
7540: U_{\eta}G^{\pm}_mU^{-1}_{\eta} &\to G^{\pm}_{m\pm \eta} \cr
7541: U_{\eta}J_mU^{-1}_{\eta} &\to J_m + \frac{\eta}{3}\delta_{m,0} \ .
7542: \end{align}
7543: The explicit form of the unitary operator $U_\eta$ can be obtained by
7544: \begin{align}
7545: U_{\eta} = e^{-i\sqrt{\frac{c}{3}}\eta\phi} \ ,
7546: \end{align}
7547: where 
7548: \begin{align}
7549: J = i\sqrt{\frac{c}{3}}\partial \phi \ .
7550: \end{align}
7551:  
7552: \subsection{Lichnerowicz obstruction and $a$-theorem}\label{Lic}
7553: In this appendix, we show that an apparent violation of the $a$-theorem for the  gauge theory living on the D3-brane at the tip of the Brieskorn-Pham singularities are avoided by the Lichnerowicz obstruction we reviewed in section \ref{sec:2-2-3}. Let us consider the Brieskorn-Pham type generalized conifolds
7554: \begin{align}
7555: x_1^{k_1} + x_2^{k_2} + k_3^{k_3} + k_4^{k_4} = 0 \ . \label{bifs}
7556: \end{align}
7557: 
7558: Assuming that the Reeb vector (conformal $U(1)_R$ charge) is given by the natural $\bc^*$ action induced by the charge vector $(1/k_1,1/k_2,1/k_3,1/k_4)$ on $(x_1,x_2,x_3,x_4)$, the volume of the associated Sasaki-Einstein manifold can be computed as
7559: \begin{align}
7560: V = \frac{\pi^3 k_1k_2k_3k_4 \left(\frac{1}{k_1}+\frac{1}{k_2}+\frac{1}{k_3}+\frac{1}{k_4}-1\right)^3}{27} \ .
7561: \end{align}
7562: The conjectured $a$-theorem states that $a$, which is inversely proportional to $V$ via $AdS-CFT$ correspondence, should be an decreasing function as we decrease $k_i$ as a relevant deformation. The condition is equivalent to
7563: \begin{align}
7564: -\frac{2}{k_1} +\frac{1}{k_2}+\frac{1}{k_3}+\frac{1}{k_4} \le 1 \ , \label{a-theo}
7565: \end{align}
7566: for $k_1$, and similarly for $k_2$, $k_3$ and $k_4$. 
7567: 
7568: On the other hand, the Lichnerowicz obstruction applied for the generalized conifold \label{bifs} is
7569: \begin{align}
7570: \sum_i \frac{1}{k_i} \le 1 + 3\omega_m \ , \label{Lich}
7571: \end{align}
7572: where $\omega_m$ is a charge of holomorphic functions on \eqref{bifs}. As a holomorphic function, we choose a monomial $x_i$ that has a charge $1/k_i$. Then the Lichnerowicz obstruction \eqref{Lich} directly yields the necessary and sufficient condition for the conjectured $a$-theorem \eqref{a-theo}. Physically speaking, this suggests that the unitarity of the SCFT is crucial for the establishment of the $a$-theorem as is the case with the two-dimensional $c$-theorem, where the unitarity is imperative for its proof.
7573: 
7574: \subsection{Boundary wavefunction from direct integration}\label{direct}
7575: We would like to compute the boundary wavefunction for the rolling D-brane in the Lorentzian two-dimensional black hole from the direct integration, which was not carried out in \cite{Nakayama:2005pk}.
7576: 
7577: We focus on the overlap between the minisuperspace wavefunction
7578: \begin{align}
7579: U^p_{\om}(\rho,t) &= -
7580: \frac{\Gamma^2(\nu_+)}{\Gamma(1-i\om)\Gamma(-ip)} e^{-i\om t} (\sinh
7581: \rho)^{-i\om} F(\nu_+,\nu^*_-;1-i\om;-\sinh^2\rho) ~,
7582: \label{U} 
7583: \end{align}
7584: and the classical trajectory
7585: \begin{eqnarray}
7586: \cosh(t) \sinh (\rho) = \sinh(\rho_0) \ .
7587: \end{eqnarray}
7588: Explicitly, we would like to evaluate the integral
7589: \begin{align}
7590: \Psi(p,\omega) &= \int_0^\infty \sinh\rho d\sinh\rho \int_{-\infty}^{\infty} dt \delta\left(\cosh(t) \sinh (\rho) - \sinh(\rho_0) \right) U^p_{\omega}(\rho,\theta) \cr
7591: & = \int_{-\infty}^{\infty} dt \frac{\sinh\rho_0}{\cosh^2t} U^p_{\omega}(\rho(\rho_0,t),t)  \ . \label{integ}
7592: \end{align}
7593: 
7594: After expanding the hypergeometric function as 
7595: \begin{align}
7596: & F(\frac{1}{2}-\frac{ip}{2}-\frac{i\omega}{2},\frac{1}{2}+\frac{ip}{2}-\frac{i\omega}{2};1-i\omega,-\frac{\sinh^2\rho_0}{\cosh^2t}) \cr
7597: =& \sum_{l=0} \frac{\Gamma(\frac{1}{2}-\frac{ip}{2}-\frac{i\omega}{2}+l)\Gamma(\frac{1}{2}+\frac{ip}{2}-\frac{i\omega}{2}+l)}{\Gamma(1-i\omega+l)} \times \cr &\times \frac{\Gamma(-i\omega)}{\Gamma(\frac{1}{2}-\frac{ip}{2}-\frac{i\omega}{2})\Gamma(\frac{1}{2}+\frac{ip}{2}-\frac{i\omega}{2})} \frac{(-)^l\sinh^{2l}\rho_0}{l!\cosh^{2l}t} \ .
7598: \end{align}
7599: the integration over $t$ is possible by the formula
7600: \begin{eqnarray}
7601: \int_{-\infty}^{\infty} (2 \cosh t)^{a-1} e^{i b t}
7602: \dd t = \frac{1}{2} B\left(\frac{1}{2}-\frac{a+ib}{2},
7603: \frac{1}{2}-\frac{a-ib}{2}\right) \equiv \frac{1}{2} \frac
7604: {\Gamma\left(\frac{1}{2}-\frac{a+ib}{2}\right)
7605: \Gamma\left(\frac{1}{2}-\frac{a-ib}{2}\right)} {\Gamma(1-a)}~, &
7606: \hspace{10cm} (\Re\, a <1~,~~ \left|\Im \, b \right| < 1- \Re\, a)~. 
7607: \end{eqnarray}
7608: 
7609: After collecting terms by using the duplication formula
7610: \begin{eqnarray}
7611: \sqrt{\pi} \Gamma(2l+2-i\omega) = 2^{2l+1-i\omega} \Gamma(2l+1-\frac{i\omega}{2})\Gamma(2l+\frac{3}{2}-\frac{i\omega}{2}) \ ,
7612: \end{eqnarray}
7613: we obtain\footnote{Presumably up to a numerical factor.}
7614: \begin{align}
7615: \Psi(p,\omega) =
7616: & \left(\frac{\sinh \rho_0}{2}\right)^{1-i\omega}\frac{\Gamma^2(\frac{1}{2}-\frac{ip}{2}-\frac{i\omega}{2})\Gamma(\frac{1}{2}+\frac{ip}{2}-\frac{i\omega}{2})}{\Gamma(-ip)\Gamma(1-\frac{i\omega}{2})\Gamma(\frac{3}{2}-\frac{i\omega}{2})}\cr &\times
7617:  \ _3F_2\left(1,\frac{1}{2}-\frac{ip}{2}-\frac{i\omega}{2},\frac{1}{2}-\frac{ip}{2}-\frac{i\omega}{2};1-\frac{i\omega}{2},\frac{3}{2}-\frac{i\omega}{2};-\frac{\sinh^2\rho_0}{2}\right)\ . \label{ans}
7618: \end{align}
7619: Here we have introduced the generalized hypergeometric function
7620: \begin{eqnarray}
7621: _3F_2(a_1,a_2,a_3;b_1,b_2;z) = \sum_{l=0} \frac{(a_1)_l (a_2)_l(a_3)_l}{(b_1)_l(b_2)_l}\frac{x^l}{l!} \ ,
7622: \end{eqnarray}
7623: where 
7624: \begin{eqnarray}
7625: (a)_l \equiv \frac{\Gamma(a+l)}{\Gamma(a)} \ .
7626: \end{eqnarray}
7627: 
7628: Let us discuss a particular limit of our boundary wavefunction \eqref{ans}. We take the limit $\rho_0 \to \infty$ to see the connection to our previous results \cite{Nakayama:2005pk}. To see this, we use the asymptotic expansion formula
7629: \begin{align}
7630: &_3F_2(a_1,a_2,a_3;b_1,b_2;z) \cr  &= \frac{\Gamma(b_1)\Gamma(b_2)}{\Gamma(a_1)\Gamma(a_2)\Gamma(a_3)} \sum_{k=1}^3 \frac{\Gamma(a_k)\prod_{j=1;j\neq k}^3(a_j-a_k)}{\prod_{j=1}^2 \Gamma(b_j-a_k)}(-z)^{-a_k}\left[1 + O(z^{-1})\right] \ .
7631: \end{align}
7632: In this limit, we see that it asymptotically approaches to our previous results
7633: \begin{align}
7634: \lim_{\rho_0\to \infty} \Psi(p,\omega) = B(\nu_+,
7635: \nu_-) \Gamma\Big(1+\frac{ip}{k}\Big) \, \left[ e^{-ip
7636: \rho_0} -  \frac{\cosh\left(\pi \frac{p-\om}{2}\right)}
7637: {\cosh\left(\pi \frac{p+\om}{2}\right)} e^{ip\rho_0 } \right] +O(e^{-\rho_0})~. 
7638: \label{falling D00}
7639: \end{align}
7640: So importantly, our previous results coincide with the direct evaluation of the overlap only in the limit $\rho_0 \to \infty$.
7641: 
7642: Another interesting limit is to take $\rho_0 \to 0$. If we further specialize in the zero energy overlap (i.e. $\omega = 0$), we have
7643: \begin{eqnarray}
7644: \lim_{\rho_0\to 0} \Psi(p,0) = -\sinh\rho_0 \frac{\Gamma^2(\frac{1}{2}-\frac{ip}{2})}{\Gamma(-ip)} \ ,
7645: \end{eqnarray}
7646: which coincides with the zero-winding sector of the D0-brane in the cigar (up to $\sinh\rho_0$), which is quite expected, given the origin of the minisuperspace calculation. It would be interesting but seem very difficult to compute the emission rate for general $\rho_0$.
7647: 
7648: 
7649: \newpage
7650: \bibliographystyle{utcaps}
7651: \bibliography{d}
7652: 
7653: 
7654: \end{document}
7655: 
7656: