hep-th0302085/kks.tex
1: %
2: % $Id: kks.tex,v 2.2 2003/04/02 23:36:27 frolov Stab $
3: %
4: 
5: %\documentstyle[preprint,tighten,prd,aps]{revtex}
6: \documentstyle[floats,epsfig,twocolumn,prd,aps]{revtex}
7: 
8: \begin{document} \draft
9: 
10: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
11: \newcommand{\be}{\begin{equation}}
12: \newcommand{\ee}{\end{equation}}
13: \newcommand{\ba}{\begin{eqnarray}}
14: \newcommand{\ea}{\end{eqnarray}}
15: \newcommand{\nn}{\nonumber\\}
16: \newcommand{\n}[1]{\label{#1}}
17: 
18: %\renewcommand\theequation{\thesection.\arabic{equation}}
19: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
20: 
21: \wideabs{%%% Remove before submission!
22: \title{Black Holes in a Compactified Spacetime}
23: \author{Andrei V. Frolov$^*$ and Valeri P. Frolov$^\dagger$}
24: \address{
25:   \medskip
26:   $^*$CITA, University of Toronto\\
27:   Toronto, ON, Canada, M5S 3H8\\
28:   {\rm E-mail: \texttt{frolov@cita.utoronto.ca}}
29: }
30: \address{
31:   \medskip
32:   $^\dagger$Theoretical Physics Institute, %\\
33:   Department of Physics, University of Alberta\\
34:   Edmonton, AB, Canada, T6G 2J1\\
35:   {\rm E-mail: \texttt{frolov@phys.ualberta.ca}}
36:   \medskip
37: }
38: \date{12 February 2003}
39: \maketitle
40: 
41: \begin{abstract}
42:   We discuss properties of a 4-dimensional Schwarzschild black hole in a
43:   spacetime where one of the spatial dimensions is compactified. As a
44:   result of the compactification the event horizon of the black hole is
45:   distorted. We use Weyl coordinates to obtain the solution describing
46:   such a distorted black hole. This solution is a special case of the
47:   Israel-Khan metric. We study the properties of the compactified
48:   Schwarzschild black hole, and develop an approximation which allows
49:   one to find the size, shape, surface gravity and other characteristics
50:   of the distorted horizon with a very high accuracy in a simple
51:   analytical form. We also discuss possible instabilities of a black
52:   hole in the compactified space.
53: \end{abstract}
54: 
55: \pacs{PACS numbers: 04.50.+h, 98.80.Cq \hfill CITA-2003-04, Alberta-Thy-03-03}
56: }
57: \narrowtext
58: 
59: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
60: 
61: 
62: \section{Introduction}\label{s1}
63: 
64: Black hole solutions in a compactified spacetime have been studied in
65: many publications. A lot of attention was paid to Kaluza-Klein
66: higher dimensional black holes. By compactifying black hole solutions
67: along Killing directions one obtains a lower dimensional solutions
68: of Einstein equations with additional scalar, vector and other fields
69: (see e.g. \cite{GiWi:86} and references therein). The generation of
70: black hole and black string solutions by the Kaluza-Klein procedure
71: was extensively used in the string theory (see e.g. \cite{Lars:00}
72: and references therein).
73: 
74: A solution which we consider in this paper is of a different nature. We
75: study a Schwarzschild black hole in a spacetime with one compactified
76: spatial dimension. This dimension does not coincide with any Killing
77: vector, for this reason the black hole metric is distorted as a result
78: of the compactification.
79: 
80: Recent interest in compactified spacetimes with black holes is
81: connected with brane-world models. General properties of black holes in
82: the Randall-Sundrum model were discussed in \cite{ChHaRe:00,EmHoMy:00}.
83: In the latter paper 4-dimensional C-metric was used to obtain an exact
84: 3+1 dimensional black hole solution in AdS spacetime with the
85: Randall-Sundrum brane. Black holes in RS braneworlds were discussed in
86: a number of publications (see e.g.~\cite{KuTaNa:03} and references
87: therein).
88: 
89: Black hole solutions in a spacetime with compactified dimensions are
90: also interesting in connection with other type of brane models, which
91: were considered historically first in \cite{BW}. In ADD-type of brane
92: worlds the tension of the brane can be not very large. If one neglects
93: its action on the gravitational field of a black hole, one obtains a
94: black hole in a spacetime where some of the dimensions are
95: compactified. Compactification of special class of solutions,
96: generalized Majumdar-Papapetrou metrics, was discussed by Myers
97: \cite{Myers:87}. In this paper he also made some general remarks
98: concerning compactification of the 4-dimensional Schwarzschild metric.
99: Some of the properties of compactified 4-dimensional Schwarzschild
100: metrics were also considered in \cite{BoPe:90}. For a recent discussion
101: of higher dimensional black holes on cylinders see \cite{HaOb:02}.
102: 
103: In this paper we study a solution describing a 4-dimensional
104: Schwarzschild black hole in a spacetime where one of the dimensions is
105: compactified. This solution is a special case of the Israel-Khan metric
106: \cite{IsKh:64}. Its general properties were discussed by Korotkin and
107: Nicolai \cite{KoNi:94}. As a result of the compactification the event
108: horizon of the black hole is distorted. In our paper we focus our
109: attention on the properties of the distorted horizon. We use Weyl
110: coordinates to obtain a solution describing such a distorted black
111: hole. This approach to study of axisymmetric static black holes is well
112: known and was developed long time ago by Geroch and Hartle \cite{GeHa:81}
113: (see also \cite{FrNo:98})\footnote{For generalization of this approach
114: to the case of electrically charged distorted 4-D black holes see
115: \cite{FaKr:01,Yaza:01}. A generalization of the Weyl method to higher
116: dimensional spacetimes was discussed in \cite{EmRe:02}. An initial
117: value problem for 5-D black holes was discussed in \cite{ShSh:00,SoPi:02}}.
118: In Weyl coordinates, the metric describing a distorted 4-dimensional
119: black hole contains 2 arbitrary functions. One of them, playing a role
120: of the gravitational potential, obeys a homogeneous linear equation.
121: Because of the linearity, one can present the solution as a linear
122: superposition of the unperturbed Schwarzschild gravitational potential,
123: and its perturbation. After this, the second function which enters the
124: solution can be obtained by simple integration. To find the
125: gravitational potential one can either use the Green's function method
126: or to expand a solution into a series over the eigenmodes. We discuss
127: both of the methods since they give two different convenient
128: representations for the solution. We develop an approximation which
129: allows one to find the size, shape, surface gravity and other
130: characteristics of the distorted horizon with very high accuracy in a
131: simple analytical form. We study properties of compactified
132: Schwarzschild black holes and discuss their possible instability.
133: 
134: The paper is organized as follows. We recall the main properties of 4-D
135: distorted black holes in Section~2. In Section~3, we obtain the
136: solution for a static 4-dimensional black hole in a spacetime with 1
137: compactified dimension. In Section~4, we study this solution. In
138: particular we discuss its asymptotic form at large distances, and the
139: size, form and shape of the horizon. We conclude the paper by general
140: remarks in Section~5.
141: 
142: 
143: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
144: 
145: \section{4-dimensional Weyl black holes}
146: 
147: \subsection{Weyl form of the Schwarzschild metric}
148: 
149: A static axisymmetric 4 dimensional metric in the canonical Weyl
150: coordinates takes the form \cite{GeHa:81,EmRe:02,FrNo:98}
151: \be\n{2.1}
152: dS^2=-e^{2U}\, dT^2 +e^{-2U}\, \left[ e^{2V}\, (dR^2 +dZ^2)+
153: R^2\, d\phi^2 \right]\, ,
154: \ee
155: where $U$ and $V$ are functions of $R$ and $Z$. This metric is a
156: solution of vacuum Einstein equations if and only if these functions
157: obey the equations
158: \be\n{2.2}
159: {\partial^2 U\over \partial R^2}+{1\over R}\,{\partial U\over \partial
160: R}+
161: {\partial^2 U\over \partial Z^2} =0\, ,
162: \ee
163: \be\n{2.3}
164: V_{,R}=R\, (U_{,R}^2-U_{,Z}^2)\, ,\hspace{1em}
165: V_{,Z}=2R\, U_{,R}\, U_{,Z}\, .
166: \ee
167: Let
168: \be\n{2.4}
169: dl^2=dR^2 +R^2\, d\psi^2+dZ^2\,
170: \ee
171: be an auxiliary 3 dimensional flat metric, then solutions of
172: (\ref{2.2}) coincide with axially symmetric solutions of the 3
173: dimensional Laplace equation
174: \be\n{2.5}
175: \Delta\, U=0\, ,
176: \ee
177: where $\Delta$ is a flat Laplace operator in the metric (\ref{2.4}).
178: It is easy to check that the equation (\ref{2.2}) plays the role of
179: the integrability condition for the linear first order equations
180: (\ref{2.3}). The regularity condition implies that at regular points
181: of the symmetry axis $R=0$
182: \be\n{2.6}
183: \lim_{R\to 0}\, V(R,Z)=0\, .
184: \ee
185: In fact, if $V(0,Z_0)=0$ at any point $Z_0$ of $Z$-axis then
186: (\ref{2.3}) implies that $V(0,Z)=0$ at any other point of the $Z$-axis
187: which is connected with $Z_0$.
188: 
189: For a four dimensional Schwarzschild metric, the function $U$ is the
190: potential of an infinitely thin finite rod of mass $1/2$ per unit
191: length located at $-M\le Z\le M$ portion of the $Z$-axis
192: \be\n{2.7}
193: \Delta\, U=4\pi j\, ,\hspace{1em} j={1\over 4\pi}{\delta(\rho)\over
194: \rho}\, \Theta(z/M)\, ,
195: \ee
196: where
197: \be\n{2.8}
198: \Theta(x)=\left\{ \begin{array}{cc}
199: 1\, , & \ \ |x|\le 1 \, ,\\
200: 0 \, ,& \ \ |x|>1 \, .
201: \end{array}
202: \right.
203: \ee
204: The corresponding solution is
205: \ba\n{2.9}
206: U_S(R,Z)
207:   &=& -{1\over 2}\int\limits_{-M}^{M}\, {dZ'\over \sqrt{R^2+(Z-Z')^2}}\, \\
208:   &=& -{1\over 2}\, \ln\left[ {\sqrt{(M-Z)^2+R^2}-Z +M \over \sqrt{(M+Z)^2+R^2}-Z-M)} \right]\, .\nonumber
209: \ea
210: The integral representation in the right hand side of equation
211: (\ref{2.9}) is obtained by using the 3-dimensional Green's function
212: for the equation (\ref{2.5}), which is of the form
213: \ba\n{2.10}
214: G^{(3)}({\bf x},{\bf x}')
215:   &=& {1\over 4\pi |{\bf x}-{\bf x}'|}\,\\
216:   && \hspace{-4em}
217:    = {1\over 4\pi}\, {1\over\sqrt{R^2+{R'}^2-2RR'\cos(\psi-\psi')+(Z-Z')^2}}\, .\nonumber
218: \ea
219: Sometimes the solution (\ref{2.9}) is presented in another equivalent
220: form
221: \be\n{2.11}
222: U_S(R,Z)={1\over 2}\ln \left({L-M\over L+M}\right)\, ,
223: \ee
224: \be\n{2.12}
225: L={1\over 2}(L_+ + L_-)\, ,\hspace{1em}
226: L_{\pm}=\sqrt{R^2+(Z\pm M)^2}\, .
227: \ee
228: The function $V_S(R,Z)$ for the Schwarzschild metric can be found
229: either by solving equations (\ref{2.3}) or by direct change of the
230: coordinates
231: \be\n{2.13}
232: R=\sqrt{r(r-2M)}\sin\theta\, ,\hspace{1em}
233: Z=(r-M)\cos\theta\, .
234: \ee
235: One has
236: \be\n{2.14}
237: V_S(R,Z)={1\over 2}\ln \left({L^2-M^2\over L^2-\eta^2}\right)\,,
238: \ee
239: \be
240: \eta={1\over 2}(L_+ -L_-)\, .
241: \ee
242: In the coordinates $(R,Z)$ the black hole horizon $H$ is the line
243: segment $-M\le Z\le M$ of the $R=0$ axis.
244: 
245: \subsection{A distorted black hole}
246: 
247: General static axisymmetric distorted black holes were studied in
248: \cite{GeHa:81}. A distorted black hole is described by a static
249: axisymmetric Weyl metric with a regular Killing horizon. One can
250: write the solution $(U,V)$ for a distorted black hole as
251: \be\n{2.15}
252: U=U_S+\hat{U}\, ,\hspace{1em} V=V_S+\hat{V}\, ,
253: \ee
254: where $(U_S,V_S)$ is the Schwarzschild solution with mass $M$. Since
255: both $V$ and $V_S$ vanish at the axis $R=0$ outside the horizon,
256: the function $\hat{V}$ has the same property. The function $\hat{U}$
257: obeys the homogeneous equation (\ref{2.2}), while the equations for
258: $\hat{V}$ follow from (\ref{2.3}). One of these equations is of the
259: form
260: \be\n{2.16}
261: \hat{V}_{,Z}=2R\left(
262: {U_S}_{,R}\,\hat{U}_{,Z} + {U_S}_{,Z}\,\hat{U}_{,R}+\hat{U}_{,R}\,\hat{U}_{,Z} \right)\, .
263: \ee
264: Near the horizon $\hat{U}$ is regular, while ${U_S}_{,R}=O( R^{-1})$
265: and ${U_S}_{,Z}=O(1)$. Thus near the horizon $\hat{V}_{,Z}\sim
266: 2\hat{U}_{,Z}$. Integrating this relation along the horizon from $Z=-M$
267: to $Z=M$ and using the relations $\hat{V}(0,-M)=\hat{V}(0,M)=0$, we
268: obtain that $\hat{U}$ has the same value $u$ at both ends of the line
269: segment $H$. By integrating the same equation along the segment $H$
270: from the end point to an arbitrary point of $H$ one obtains for $-M\le
271: Z\le M$
272: \be\n{2.17}
273: \hat{V}(0,Z)=2 \left[\hat{U}(0,Z)-u\right]\, .
274: \ee
275: Geroch and Hartle \cite{GeHa:81} demonstrated that if $\hat{U}$ is a
276: regular smooth solution of (\ref{2.5}) in any small open neighborhood
277: of $H$ (including $H$ itself) which takes the same values, $u$, on
278: the both ends of the segment $H$, then the solution is regular at the
279: horizon and describes a distorted black hole.
280: 
281: Using the coordinate transformation
282: \ba\n{2.18}
283: R&=&e^u\, \sqrt{r(r-2M_0)}\, \sin\theta\, ,\\
284: Z&=&e^u\, (r-M_0)\, \cos\theta\, ,\nonumber
285: \ea
286: and defining
287: \be\n{2.19}
288: M_0=M\, e^{-u}\,,
289: \ee
290: it is possible to recast the metric (\ref{2.1}) of a distorted black
291: hole into the form
292: \ba\n{2.20}
293: dS^2 &=& -e^{-2\hat{U}}\, \left(1-{2M_0\over r}\right)dT^2 \\
294:      && +e^{2(\hat{V}-\hat{U}+u)}\, \left(1-{2M_0\over r}\right)^{-1}\, dr^2 \nn
295:      && +e^{2(\hat{V}-\hat{U}+u)}\, r^2\, \left(d\theta^2+e^{-2\hat{V}}\, \sin^2\theta\, d\phi^2\right)\, .\nonumber
296: \ea
297: In these coordinates, the event horizon is described by the equation
298: $r=2M_0$, and the 2-dimensional metric on its surface is
299: \be\n{2.21}
300: d\gamma^2
301: =4M_0^2\, \left[ e^{2(\hat{U}-u)}\, d\theta^2
302: +e^{-2(\hat{U}-u)}\, \sin^2\theta\, d\phi^2 \right]\, .
303: \ee
304: The horizon surface has area
305: \be\n{2.22}
306: A=16\pi M_0^2\, .
307: \ee
308: It is a sphere deformed in an axisymmetric manner. The surface gravity
309: $\kappa$ is constant over the horizon surface:
310: \be\n{2.23}
311: \kappa={e^u\over 4M_0}\, .
312: \ee
313: 
314: 
315: 
316: \section{4-D compactified Schwarzschild black hole}
317: 
318: \subsection{Compactified Weyl metric}
319: 
320: In what follows it is convenient to rewrite the Weyl metric
321: (\ref{2.1}) in the dimensionless form $dS^2=L^2\, ds^2$,
322: \be\n{3.1}
323: ds^2=-e^{2U}\, dt^2 +e^{-2U}\, \left[ e^{2V}\, (d\rho^2 +dz^2)+
324: \rho^2\, d\phi^2 \right]\, ,
325: \ee
326: where $L$ is the scale parameter of the dimensionality of the length
327: and
328: \be\n{3.2}
329: t={T\over L}\, ,\quad \rho={R\over L}\, ,\quad z={Z\over L}\, .
330: \ee
331: are dimensionless coordinates. We shall also use instead of mass $M$
332: its dimensionless version $\mu=M/L$. The Schwarzschild solution
333: (\ref{2.9}) then can be rewritten as
334: \be\n{3.3}
335: U_S(\rho,z)= -{1\over 2}\, \log\left[ {\sqrt{
336: (\mu-z)^2+\rho^2}-z +\mu \over \sqrt{
337: (\mu+z)^2+\rho^2}-z-\mu)} \right]\, .
338: \ee
339: 
340: For $|z|>\mu$, the gravitational potential $U_S$ remains finite at the
341: symmetry axis
342: \be\n{3.4}
343: U_S(0,z)= {1\over 2}\,\ln {z-\mu\over z+\mu}\, ,\hspace{0.5cm}
344: |z|>\mu\, .
345: \ee
346: For $|z|\le \mu$, the gravitational potential $U_S$ is divergent at
347: $\rho=0$. The leading divergent term is
348: \be\n{3.5}
349: U_S(\rho,z)\sim {1\over 2}\,\ln {\rho^2\over 4(\mu^2-z^2)}\, ,\hspace{0.5cm}
350: |z|\le \mu\, .
351: \ee
352: 
353: We will now obtain a new solution describing a Schwarz\-schild black
354: hole in a space in which $Z$-coordinate is compactified. We will call
355: this solution a compactified Schwarzschild metric, or briefly
356: CS-metric. For this purpose we assume that the coordinate $Z$ is
357: periodic with a period $2\pi L$. We shall use the radius of
358: compactification $L$ as the scale factor.
359: 
360: Our space manifold ${\cal M}$ has topology $S^1\times R^2$ and we are
361: looking for a solution of the equation (\ref{2.5}) on ${\cal M}$ which
362: is periodic in $z$ with the period $2\pi$, \ $z\in (-\pi,\pi)$. The
363: source for this solution is an infinitely thin rod of the linear
364: density $1/2$ located along $z$ axis in the interval $(-\mu,\mu)$, \,
365: $\mu\le \pi$. This problem can be solved by two different methods,
366: either by using Green's functions or by expanding a solution into a
367: series over the eigenmodes. We discuss both of the methods since they
368: give two different convenient representations for the solution. We
369: begin with the method of Green's functions.
370: 
371: \subsection{3-D Green's function}
372: 
373: To obtain this solution we proceed as follows. Our first step is to
374: obtain a 3-dimensional Green's function $G^{(3)}_{\cal M}$ on the
375: manifold ${\cal M}$. It can be done, for example, by the method of
376: images applied to the Green's function for equation (\ref{2.5}) which
377: gives the series representation for $G^{(3)}_{\cal M}$. It is more
378: convenient to use another method which gives the integral
379: representation. For this purpose we note that the flat 3-dimensional
380: Green's function can be obtained by the dimensional reduction from the
381: 4-dimensional one. Namely let ${\bf X}=(X,Y,Z,W)$
382: \be\n{3.6}
383: dh^2=d{\bf X}^2=dX^2+dY^2+dZ^2+dW^2\, ,
384: \ee
385: then
386: \be\n{3.7}
387: G^{(3)}({\bf x},{\bf x}')\equiv {1\over 4\pi |{\bf x}-{\bf x}'|}
388: =\int\limits_{-\infty}^{\infty} \, dW \, G^{(4)}({\bf X},{\bf X}')\, ,
389: \ee
390: where ${\bf x}=(X,Y,Z)$,
391: \be\n{3.8}
392: G^{(4)}({\bf X},{\bf X}')={1\over 4\pi^2}{1\over |{\bf X}-{\bf X}'|^2}\, ,
393: \ee
394: and $G^{(4)}(X,X')$ is the Green's function for the Laplace
395: operator
396: \be\n{3.9}
397: \Delta^{(4)}\, G^{(4)}({\bf X},{\bf X}')=-\delta^4({\bf X}-{\bf X}')\, ,
398: \ee
399: 
400: Denote
401: \be\n{3.10}
402: G^{(4)}_{\cal M}({\bf X},{\bf X}')= {1\over 4\pi^2}\, \sum_{n=-\infty}^{\infty}\,
403: {1\over (Z-Z'+2\pi L n)^2+ B^2}\, .
404: \ee
405: where
406: \be\n{3.11}
407: B^2=(X-X')^2+(Y-Y')^2+(W-W')^2\, .
408: \ee
409: The function $G^{(4)}_{\cal M}$ is periodic in $Z$ with the period
410: $2\pi L$ and is a Green's function on the manifold ${\cal M}$. The sum
411: can be calculated explicitly by using the following relation
412: \be\n{3.12}
413: \sum_{-\infty}^{\infty}{1\over (a+n)^2+b^2}={\pi\over b}{\sinh(2\pi
414: b)\over \cosh(2\pi b)-\cos(2\pi a)}\, .
415: \ee
416: Thus one has
417: \be\n{3.13}
418: G^{(4)}_{\cal M}({\bf X},{\bf X}')= {1\over 8\pi^2 L^2 \beta}\,
419: {\sinh \beta\over [\cosh\beta-\cos(z-z')]}\, ,
420: \ee
421: where $\beta =B/L$. This Green's function has a pole at $\beta=z-z'=0$,
422: that is when the points ${\bf X}$ and ${\bf X}'$ coincide. At far
423: distance, $\beta\gg L$, this Green's function has asymptotic
424: \be\n{3.14}
425: G^{(4)}_{\cal M}({\bf X},{\bf X}')\sim {1\over 8\pi^2 L^2 \beta}\, ,
426: \ee
427: and hence it behaves as if the space had one dimension less. It
428: is obviously a result of compactification.
429: 
430: In the reduction procedure this creates a technical problem since the
431: integral over $w$ becomes divergent. It is easy to deal with this
432: problem as follows. Denote
433: \be\n{3.15}
434: G^{(4,\alpha)}_{\cal M}({\bf X},{\bf X}')= G^{(4,\text{reg})}_{\cal M}({\bf X},{\bf X}')
435: +{1\over 8\pi^2 L^2 }{1\over (\beta^2+b^2)^{\alpha/2}}\, ,
436: \ee
437: \ba\n{3.16}
438: G^{(4,\text{reg})}_{\cal M}({\bf X},{\bf X}') = \hspace{-6em} &&\\
439: && = {1\over 8\pi^2 L^2 } \left[{1\over \beta}\,
440: \,{\sinh \beta\over \cosh\beta-\cos(z-z')}
441: -{1\over \sqrt{\beta^2+b^2}}\right]\, . \nonumber
442: \ea
443: Here $b$ is any positive number. For $\alpha=1$, $G^{(4,\alpha)}_{\cal
444: M}$ does not depend on $b$ and coincides with (\ref{3.7}). At large
445: $\beta$ the term $G^{(4,\text{reg})}_{\cal M}$ has asymptotic behavior
446: $\sim \beta^{-2}$.
447: 
448: We also have
449: \ba\n{3.17}
450: \int\limits_{-\infty}^{\infty}\, {dw\over (\beta^2+b^2)^{\alpha/2}} = \hspace{-7em} && \\
451:  && = {1\over 2\sqrt{\pi}}{ [\sigma^2+b^2]^{(1-\alpha)/2} \Gamma( (\alpha-1)/2)\over \Gamma(\alpha/2)}\, \nn
452:  && \sim {1\over \pi}\, \left[ {1\over \alpha-1}+\ln 2 -{1\over 2}\ln (\sigma^2+b^2) \right] +O(\alpha-1) \nonumber\, .
453: \ea
454: Here $\sigma^2=(x-x')^2+(y-y')^2$. By omitting unimportant (divergent)
455: constant we regularize the expression for the integral.
456: 
457: By using the reduction procedure (\ref{3.7}) we get
458: \ba\n{3.18}
459: G^{(3)}_{\cal M}({\bf x},{\bf x}') &=& \int\limits_{-\infty}^{\infty}\, dW\,
460: G^{(4,\text{reg})}_{\cal M}({\bf X},{\bf X}') \\
461:  &&\hspace{1.5em} -{1\over 16\pi^2 L}\, \ln (\sigma^2+b^2)\, . \nonumber
462: \ea
463: 
464: \subsection{Integral representation for the gravitational potential}
465: 
466: To obtain the potential $U(\rho,z)$ which determines the black hole
467: metric we need to integrate $G^{(3)}_{\cal M}({\bf x},{\bf x}')$ with
468: respect to ${\bf x}'$ along the interval $(-M,M)$ at $R'=0$ axis. It is
469: convenient to use the representation (\ref{3.18}) and to change the
470: order of integrals. We use the following integral $(a>1, 0<\mu<\pi,
471: -\pi< z <\pi)$
472: \ba\n{3.19}
473:   \int\limits_{-\mu}^{\mu}\, {dz'\over a-\cos(z'-z)} = \hspace{-8em} &&\\
474:   && = {2\over \sqrt{a^2-1}}\left\{\textstyle
475:     \arctan\left[p\tan\left({\mu+z\over 2}\right)\right]
476:     +\pi\vartheta(\mu+z-\pi) \right.\nn
477:   &&\left.\textstyle \hspace{4.5em} +
478:     \arctan\left[p\tan\left({\mu-z\over 2}\right)\right]
479:     +\pi\vartheta(\mu-z-\pi)\right\} \, ,\nonumber
480: \ea
481: where $p=\sqrt{a+1\over a-1}$. We understand $\arctan$ to be the
482: principal value and include $\vartheta$-functions to get the correct
483: value over all the interval $-\pi <z< \pi$. We also change the
484: parameter of integration $W$ to $w=W/L$ and take into account that the
485: integrand is an even function of $w$. After these manipulations we
486: obtain
487: \ba\n{3.20}
488: U(\rho,z) &=& -{1\over \pi}
489: \int\limits_0^{\infty}\, dw\left(
490: {{\cal U}(\beta,z)\over \beta}-{\mu\over
491: \sqrt{\beta^2+b^2}}
492: \right)\\
493:  && \hspace{4em} +{\mu\over 2\pi}\ln(\rho^2+b^2)\, ,\nonumber
494: \ea
495: where
496: \ba\n{3.21}
497: {\cal U}(\beta,z)&=&{\cal V}(\beta,z) + {\cal V}(\beta,-z)\, ,\\
498: \n{3.22}
499: {\cal V}(\beta,z)&=&\textstyle
500: \arctan\left[
501: {\cosh\beta+1\over \sinh\beta}\,
502: \tan\left({\mu+z\over 2}\right)
503: \right]+\pi\vartheta(\mu+z-\pi)\, .\hspace{-1em}\nonumber
504: \ea
505: Note that now $\beta$ which enters equations (\ref{3.20}) and
506: (\ref{3.21}) is
507: \be\n{3.23}
508: \beta=\sqrt{w^2+\rho^2}\, .
509: \ee
510: 
511: A representation similar to (\ref{3.20}) can be written for the
512: Schwarzschild potential $U_S$
513: \be\n{3.24}
514: U_S(\rho,z)=-{1\over \pi}
515: \int\limits_0^{\infty}\, dw\ 
516: {{\cal U}_S(\beta,z)\over \beta}\, ,
517: \ee
518: where
519: \ba\n{3.25}
520: {\cal U}_S(\beta,z)&=&{\cal V}_S(\beta,z)
521:  +{\cal V}_S(\beta,-z)\, ,\\
522: {\cal V}_S(\beta,z)&=&\textstyle
523: \arctan\left({\mu+z\over \beta}\right)\, .\nonumber
524: \ea
525: One can check that this integral really gives expression (\ref{3.3}).
526: 
527: Using these representations we obtain the following expression for the
528: quantity $\hat{U}(z)=U(0,z)-U_S(0,z)$ which determines the properties of
529: the event horizon
530: \be\n{3.26}
531: \hat{U}(z)=-{1\over \pi}
532: \int\limits_0^{\infty}\, dw\left(
533: {{\cal U}(w,z)-{\cal U}_S(w,z)
534: \over w}-{\mu\over \sqrt{w^2+1}}
535: \right)\, .
536: \ee
537: To obtain the redshift factor $u$ it is sufficient to calculate
538: $\hat{U}(z)$ for $z=\mu$
539: \be\n{3.27}
540: u=\hat{U}(\mu)\, .
541: \ee
542: 
543: \subsection{Series representation for the gravitational potential}
544: 
545: For numerical calculations of the gravitational potential $U$ and study
546: of its asymptotics near the black hole horizon it is convenient to use
547: another representation for $U$, namely its Fourier decomposition with
548: respect to the periodic variable $z$. Note that a function
549: $\Theta(z/\mu)$ which enters the source term (see equations
550: (\ref{2.7}--\ref{2.8})) allows the following Fourier decomposition
551: on the circle
552: \be\n{3.28}
553: \Theta(z/\mu)=a_0+\sum_{k=1}^{\infty}\, a_k\ \cos(kz)\, ,
554: \ee
555: where
556: \be\n{3.29}
557: a_0={\mu\over \pi}\, ,\hspace{0.5cm} a_k={2\over \pi k}\sin(k \mu)\, .
558: \ee
559: Using the Fourier decomposition for $U$
560: \be\n{3.30}
561: U(\rho,z)=U_0(\rho)+\sum_{k=1}^{\infty} U_k(\rho)\, \cos(kz)\, ,
562: \ee
563: we obtain the following equations for the radial functions $U_k(\rho)$
564: \be\n{3.31}
565: {d^2 U_k\over d\rho^2}+{1\over \rho}\, {d U_k\over d\rho} -k^2\, U_k=
566: a_k\, {\delta(\rho)\over \rho}\, .
567: \ee
568: For $k>0$ the solutions of these equations which are decreasing at
569: infinity are
570: \be\n{3.32}
571: U_k(\rho)=-a_k K_0(k\rho)\, ,
572: \ee
573: where $K_{\nu}(z)$ is MacDonald function. For $k=0$ the solution is
574: \be\n{3.33}
575: U_0(\rho)=a_0 \ln(\rho)\, .
576: \ee
577: Thus the gravitational potential $U$ allows the following series
578: representation
579: \be\n{3.34}
580: U(\rho, z)={\mu\over \pi}\, \ln \rho
581: -2\sum_{k=1}^{\infty}\, {\sin(k\mu)\over \pi k}\, \cos(kz)\, K_0(k\rho)
582: \, .
583: \ee
584: 
585: This representation is very convenient for studying the asymptotics
586: of the gravitational potential near the horizon. For small $\rho$ one
587: has
588: \be\n{3.35}
589: -K_0(k\rho)\sim \ln\left({k\rho\over 2} \right)+\gamma \, ,
590: \ee
591: where $\gamma \approx 0.57721$ is Euler's constant. Substituting these
592: asymptotics into (\ref{3.34}) and combining the terms one obtains
593: \ba\n{3.36}
594: U(\rho,z) &\sim&
595:   \left[ \ln{\rho\over 2}\,+\gamma \right]\, \Theta(z/\mu)
596:   -{\mu\over \pi}\,\left(\ln {1\over 2}\,+\gamma \right)\, \\
597:   &&
598:   +{1\over \pi}\, \sum_{k=1}^{\infty}\, {\ln k\over k}\,
599:   \Big[\sin(k(\mu+z))+\sin(k(\mu-z))\Big]\, .\nonumber
600: \ea
601: Using the relation (see equation 5.5.1.24 in \cite{PBM})
602: \ba\n{3.37}
603: \sum_{k=1}^{\infty}\, {\ln k\over k}\,\sin(kx) &=&
604:   {x-\pi\over 2}\, (\gamma+\ln 2\pi) \\
605:   &&\hspace{-2em} + {\pi\over 2}\, \ln \left| {1\over \pi}\sin {x \over 2}\,
606:     \Gamma^2\left({x\over 2\pi}\right) \right| \nonumber
607: \ea
608: valid for $0 \le x < 2\pi$, one gets for $|z|\le \mu$
609: \ba\n{3.38}
610: U(\rho,z) &\sim& \ln {\rho\over\pi}\, + {\mu\over\pi}\, \ln(4\pi) \\
611:   && +{1\over 2}\ln \left|{1\over \pi}\, \sin{\mu+z\over 2}\,
612:     \Gamma^2\left({\mu+z\over 2\pi}\right) \right| \nn
613:   && +{1\over 2}\ln \left|{1\over \pi}\, \sin{\mu-z\over 2}\,
614:     \Gamma^2\left({\mu-z\over 2\pi}\right) \right|\, .\nonumber
615: \ea
616: Using asymptotic (\ref{3.5}) of the Schwarzschild potential $U_S$ near horizon,
617: one can present $\hat{U}(z)=\lim\limits_{\rho\to 0}[U(\rho,z)-U_S(\rho,z)]$
618: in the region $|z|\le \mu$ in the following form
619: \be\n{3.39}
620: \hat{U}(z) = {\mu\over \pi}\, \ln(4\pi) + {1\over 2}\, \ln \left[\textstyle
621: f({\mu+z\over 2})f({\mu-z\over 2})\right]\, ,
622: \ee
623: where the function $f(x)$ is defined by
624: \be\n{3.40}
625: f(x)={1\over \pi^2}\, x\, \sin x\, \Gamma^2(\textstyle{x\over\pi})\, .
626: \ee
627: It has the following properties:
628: \be\n{3.41}
629: f(0)=1\, ,\hspace{1em}f(\textstyle{\pi \over 2})={1 \over 2} \, ,\hspace{1em}f(\pi)=0\, .
630: \ee
631: In fact, in the interval $0\le x \le \pi$ it can be approximated by a
632: linear function
633: \be\n{3.42}
634: f(x)\approx 1-{x\over \pi}\,
635: \ee
636: with an accuracy of order of 1\% .
637: 
638: Making similar calculations for $|z|\ge \mu$ one obtains
639: \be\n{3.38a}
640: U(0,z)= {\mu\over \pi}\, \ln (4\pi)
641: +{1\over 2}\ln \left[ { f({|z|+\mu\over 2})\over f({|z|-\mu\over
642: 2})}\right]+{1\over 2}\ln {|z|-\mu \over |z|+\mu}\, .
643: \ee
644: An approximate value of $U(0,z)$ in this region is
645: \be\n{3.38b}
646: U(0,z)\approx {\mu\over \pi}\, \ln (4\pi)
647: +{1\over 2}\ln \left[ {[2\pi-(|z|+\mu)](|z|-\mu)\over
648: [2\pi-(|z|-\mu)](|z|+\mu)}\right]\, .
649: \ee
650: 
651: \subsection{Solutions}
652: 
653: \begin{figure*}
654: \begin{center}\begin{tabular}{c@{\hspace{1cm}}c}
655:   \epsfig{file=surface-U05.eps, width=8.5cm} &
656:   \epsfig{file=surface-U20.eps, width=8.5cm} \\
657:   \\ \\
658:   \epsfig{file=surface-V05.eps, width=8.5cm} &
659:   \epsfig{file=surface-V20.eps, width=8.5cm} \\
660: \end{tabular}\end{center}
661: \caption{
662:   Compactified Schwarzschild black hole solutions for $\mu=0.5$ (left)
663:   and $\mu=2.0$ (right). The surface plots show the gravitational
664:   potential $U(\rho,z)$ (top) and the function $V(\rho,z)$ (bottom);
665:   red and blue contours represent equipotential surfaces of $U$ and $V$
666:   correspondingly.
667: }
668: \label{equipot}
669: \end{figure*}
670: 
671: To find the gravitational potential $U(\rho,z)$ one can use either its
672: integral representation (\ref{3.20}) or the series (\ref{3.34}). We
673: used both methods. Integrals (\ref{3.20}) were evaluated using Maple,
674: while the series (\ref{3.34}) were implemented in C code using FFT
675: techniques. Both methods give results which agree with high accuracy,
676: but of course the C implementation is much more computationally
677: efficient. The function $V(\rho,z)$ was recovered by direct integration
678: of differential equation (\ref{2.3}) by finite differencing in
679: $Z$-direction. The gravitational potential $U(\rho,z)$, function
680: $V(\rho,z)$, and their equipotential surfaces for two different values
681: of $\mu$ are shown in Figure~\ref{equipot}.
682: 
683: 
684: 
685: \section{Properties of CS black holes}
686: 
687: \begin{figure*}
688:   \centerline{
689:     \epsfig{file=uu_mu.eps, width=3in}\hspace{1cm}
690:     \epsfig{file=mu0.eps, width=3in}
691:   }
692:   \caption{
693:     Redshift factor $u$ (left) and the irreducible mass $\mu_0=\mu\, \exp(-u)$ as functions of $\mu$.
694:   }
695:   \label{u}
696: \end{figure*}
697: 
698: \subsection{Large distance asymptotics}
699: 
700: Let us first analyze the asymptotic behavior of the CS-metric at large
701: distance $\rho$. For this purpose we use the integral representation
702: (\ref{3.20}) for $U$. It is easy to check that the integrand
703: expression at large $\rho$ is of order of $O(\beta^{-2})$ and hence
704: the integral is of order of $\rho^{-1}$. Thus the ln-term in the square
705: brackets in (\ref{3.20}) is leading at the infinity so that
706: \be\n{4.1}
707: U(\rho,z)|_{\rho\to\infty}\sim {\mu\over \pi}\ln\rho\, .
708: \ee\n{4.2}
709: Using equation (\ref{2.3}) we also get
710: \be\n{4.3}
711: V(\rho,z)|_{\rho\to\infty}\sim {\mu^2\over \pi^2}\ln\rho\, .
712: \ee
713: The metric (\ref{3.1}) in the asymptotic region $\rho\to\infty$ is of the form
714: \be\n{4.4}
715: ds^2=-\rho^{2{\mu\over\pi}}\, dt^2+\rho^{-2{\mu\over\pi}(1-{\mu\over\pi})}\,
716: (d\rho^2+dz^2)+ \rho^{-2{\mu\over\pi}}\, \rho^2\, d\phi^2 \, .
717: \ee
718: The proper size of a closed Killing trajectory for the vector
719: $\partial_z$ is
720: \be\n{4.4a}
721: C_z=2\pi L\rho^{-{\mu\over \pi}(1-{\mu\over \pi})}\, .
722: \ee
723: 
724: The metric (\ref{4.4}) coincides with the special case $(a_1=a_2)$ of
725: the Kasner solution \cite{Kasner}
726: \ba\n{4.5}
727: &ds^2=-\rho^{2a_0}\, dt^2+ \rho^{2a_1}\, d\rho^2+ \rho^{2a_2}\, dz^2
728: +\rho^{2a_3}\, d\phi^2\, ,&\\
729: \n{4.6}
730: &a_1+1=a_2+a_3+a_0\,,&\nn
731: &(a_1+1)^2=a_2^2+a_3^2+a_0^2\, .&\nonumber
732: \ea
733: 
734: One can rewrite the metric (\ref{4.3}) by using the proper-distance
735: coordinate $l$. For small $\mu$
736: \be\n{4.7}
737: l={\rho^{1-{\mu\over\pi}}\over 1-{\mu\over\pi}}\, ,
738: \ee
739: and the metric in the $(\rho,\phi)$-sector takes the form
740: \be\n{4.8}
741: dl^2+\left(1-{\mu\over\pi}\right)^2\, l^2\, d\phi^2\, .
742: \ee
743: Thus the metric of the CS black hole has an angle deficit $2\mu$ at
744: infinity.
745: 
746: The asymptotic form of the metric can be used to determine the mass of
747: the system. Let $\xi_{(t)}^{\mu}$ be a timelike Killing vector and
748: $\Sigma$ be a 2-D surface lying inside $t$=const hypersurface, then
749: the Komar mass $m$ is defined as
750: \be\n{4.8a}
751: m={1\over 4\pi}\, \int_{\Sigma}\, \xi_{(t)}^{\mu;\nu}\, d\sigma_{\mu\nu}\, .
752: \ee
753: For simplicity we choose $\Sigma$ so that $t$=const and
754: $\rho=\rho_0$=const. For this choice
755: \ba\n{4.8b}
756: d\sigma_{\mu\nu} &=& {1\over 2}\, \delta^{0}_{[\mu}\,
757: \delta^{1}_{\nu]}\,\rho_0^{1+2a_1}\, dz\, d\phi\, ,\\
758: \n{4.8c}
759: \xi_{\mu;\nu} &=& -2a_0\, \rho_0^{2a_0-1}\, \delta_{[\mu}^{0}\,
760: \delta_{\nu]}^{1}\, ,\nn
761: \xi^{\mu;\nu} &=& 2a_0\, \rho_0^{-2a_1-1}\, \delta^{[\mu}_{0}\,
762: \delta^{\nu]}_{1}\, .\nonumber
763: \ea
764: Substituting these expressions into (\ref{4.8a}) and taking the
765: integral we get $m=\mu$. Since all our quantities are normalized by the
766: radius of compactification $L$, we obtain that the Komar mass of our
767: system is $M=L\mu$.
768: 
769: \subsection{Redshift factor, surface gravity, and proper distance
770: between black hole poles}
771: 
772: Using equation (\ref{3.39}), we obtain for the redshift factor $u$ the
773: following expression
774: \be\n{3.43}
775: u={\mu\over \pi}\ln (4\pi)+{1\over 2}\ln f(\mu)\, .
776: \ee
777: Figure~\ref{u}(left) shows dependence of the redshift factor $u$ on
778: parameter $\mu$. Using the approximation (\ref{3.42}) we can write
779: \be\n{3.44}
780: u\approx {\mu\over \pi}\, \ln(4\pi)+{1\over 2}\ln \left(1-{\mu\over \pi}\right)\, .
781: \ee
782: The redshift factor $u$ has maximum $u_*$
783: \be\n{3.44a}
784: u_{*}=\ln(4\pi)-{1\over 2}\left[1+\ln 2 +\ln(\ln (4\pi)) \right] \approx 1.22
785: \ee
786: at
787: \be\n{3.44b}
788: \mu=\pi(1-1/(2\ln(4\pi)))\approx 2.52\, .
789: \ee
790: For $\mu>\mu_*$ the function $u$ rapidly falls down, becoming negative
791: and logarithmically divergent at $\mu=\pi$.
792: 
793: In the same approximation we get the following expressions for the
794: irreducible mass $\mu_0$ and the surface gravity $\kappa$
795: \ba\n{3.45}
796: \mu_0=\mu\, \exp(-u) &\approx&
797:   \mu\, (4\pi)^{-{\mu\over\pi}}\, \left(1-{\mu\over\pi}\right)^{-{1\over 2}}, \\
798: \n{3.46}
799: \kappa={e^{-2u}\over 4\mu} &\approx& {1\over 4\mu}\,
800: (2\pi)^{2{\mu\over\pi}}\left(1-{\mu\over \pi}\right)\, .
801: \ea
802: For $\mu\to\pi$, they behave as $\mu_0\to \infty$ and $\kappa\to 0$.
803: Figure~\ref{u}(right) shows the irreducible mass $\mu_0$ as
804: a function of $\mu$.
805: 
806: 
807: \begin{figure}[t]
808:   \centerline{\epsfig{file=polength.eps, width=3in}}
809:   \caption{$l/(2\pi)$ as a function of $\mu$.}
810:   \label{polength}
811: \end{figure}
812: 
813: Another invariant characteristic of the solution is the proper distance
814: between the `north pole', $z=\mu$, and `south pole', $z=-\mu$, along a
815: geodesic connecting these poles and lying outside the black hole. This
816: distance, $l(\mu)$, is
817: \ba\n{3.45a}
818: l(\mu)
819:   &=& 2\int\limits_{\mu}^{\pi}\, dz\, e^{-U(0,z)}\\
820:   &\approx& 2(4\pi)^{-\mu/\pi}\, \int\limits_{\mu}^{\pi}\, dz\, \sqrt{{(z+\mu)(2\pi-z+\mu)\over (z-\mu)(2\pi-z-\mu)}}\nn
821:   &=&\textstyle
822:    2\sqrt{\pi^2-\mu^2}\, E(\varphi,k)\,
823:    +2\mu \sqrt{ {\pi+\mu\over \pi-\mu}}\, F(\varphi,k)\,
824:    -(\pi-\mu)\, ,\hspace{-1em}\nonumber
825: \ea
826: where
827: \be\n{3.45b}
828: \varphi =\sqrt{1-\mu/\pi}\, ,\hspace{1em}k={1\over \sqrt{1-(\mu/\pi)^2}}\, .
829: \ee
830: Here $F(\varphi,k)$ and $E(\varphi,k)$ are the elliptic integrals of
831: the first and second kind, respectively. In particular one has
832: \be\n{3.45c}
833: l(0)=2\pi \, ,\hspace{1em} l(\pi)=\pi/2\, .
834: \ee
835: Figure~\ref{polength} shows $l/(2\pi)$ as a function of $\mu$.
836: It might be surprising that in the limit $\mu\to\pi$, when the
837: coordinate distance $\Delta z$ between the poles tends to 0, the
838: proper distance between them remains finite. This happens because in
839: the same limit the surface gravity tends to 0.
840: 
841: \subsection{Size and shape of the event horizon}
842: 
843: \begin{figure*}
844:   \centerline{
845:     \epsfig{file=EF.eps, height=2in}\hspace{1cm}
846:     \epsfig{file=gauss.eps, height=2in}
847:   }
848:   \caption{
849:     The shape function $\exp({\cal F}(z))$ (left) and the Gaussian
850:     curvature of the horizon $K(z)$ (right) for different values of $\mu$.
851:   }
852:   \label{shape}
853:   \label{gauss}
854: \end{figure*}
855: 
856: \begin{figure*}
857:   \medskip
858:   \centerline{\epsfig{file=embedding.eps, width=6in}}\medskip
859:   \caption{
860:     Embedding diagrams for the surface of the black hole horizon. By
861:     rotating a curve from a family shown at the plot around a horizontal
862:     axis one obtains surface isometric to the surface of a black hole
863:     described by the metric $d\sigma^2$. Different curves correspond to
864:     different values of $\mu$. The larger $\mu$ the more oblate is the
865:     form of the curve.
866:   }
867:   \label{embed}
868: \end{figure*}
869: 
870: The surface area of the distorted horizon (\ref{2.22}) written in
871: units $L^2$ is
872: \be\n{4.9}
873: A=16\pi \mu_0^2 \, ,
874: \ee
875: where $\mu_0$ is the irreducible mass (\ref{3.45}). The shape of the
876: horizon is determined by the {\em shape function}
877: \be\n{4.10}
878: {\cal F}(z)=\hat{U}(z)-u\, .
879: \ee
880: Figure~\ref{shape}(left) shows a plot of $\exp({\cal F}(z))$ for
881: several values of $\mu$. By multiplying the 2-metric on the horizon
882: $d\gamma^2$ by $(2\mu_0)^{-2}$ one obtains the metric of the 2-surface
883: which has the topology of a sphere $S^2$ and the surface area $4\pi$.
884: The metric describing this distorted sphere is
885: \be\n{4.11}
886: d\sigma^2=e^{2{\cal F}}{dz^2\over \mu^2-z^2}+e^{-2{\cal F}}
887: (\mu^2-z^2){d\phi^2\over \mu^2}\, .
888: \ee
889: 
890: The Gaussian curvature of the metric $d\sigma^2$ is $K={1\over 2}R$,
891: where $R$ is the Ricci scalar curvature. It is given by the following
892: expression
893: \be\n{4.12}
894: K =
895: e^{-2{\cal F}(z)}\, \left[1+ (\mu^2-z^2)\,[{\cal F}'' -2({\cal F}')^2
896: ]\,
897: -4z {\cal F}' \right]\, .
898: \ee
899: The Gauss-Bonnet formula gives
900: \be\n{4.13}
901: \int d^2x\, \sqrt{\sigma}\, K=4\pi\, .
902: \ee
903: For the unperturbed black hole $K=1$. As a result of deformation, the
904: CS-black hole has $K>1$ at the poles, $z=\pm \mu$, and $K<1$ at the
905: `equatorial plane', $z=0$. Figure~\ref{gauss}(right), which shows $K(z)$
906: for different values of $\mu$, illustrates this feature. This kind of
907: behavior can be easily understood as a result of self-attraction of the
908: black hole because of the compactification of the coordinate $z$.
909: 
910: Using approximation (\ref{3.42}) allows one to obtain simple analytical
911: expressions for the shape function and the Gaussian curvature. Equations
912: (\ref{3.39}) and (\ref{3.43}) give
913: \be\n{4.14}
914: {\cal F}={1\over 2}\, \ln\left[{f({\mu+z\over 2}) f({\mu-z\over 2})
915: \over f(\mu)} \right]\approx {1\over 2}\, \ln\left[ 1+{\mu^2-z^2\over
916: 4\pi(\pi-\mu)}\right]\, .
917: \ee
918: Let us write the metric $d\sigma^2$ in the form
919: \be\n{4.15}
920: d\sigma^2=F(z)\, dz^2+{d\phi^2\over \mu^2\, F(z)}\, ,
921: \ee
922: then in this approximation one has
923: \be\n{4.16}
924: F(z)\approx{1\over \mu^2-z^2}+{1\over 4\pi^2(1-\mu/\pi)}\, .
925: \ee
926: while the Gaussian curvature is
927: \be\n{4.17}
928: K\approx{16\pi^2(\pi-\mu)^2[ (2\pi-\mu)^2+3z^2]\over [(2\pi-\mu)^2-z^2]^3}\, .
929: \ee
930: The Gaussian curvature is positive in the interval $|z|<\mu$.
931: 
932: It is interesting to note that the horizon geometry of the CS-black hole
933: coincides (up to a constant factor) with the geometry on the 2-D
934: surface of the horizon of the Euclidean 4-D Kerr black hole. This fact
935: can be easily checked since the induced 2-D geometry of the horizon of
936: the Kerr black hole is (see e.g. equation 3.5.4 in \cite{FrNo:98})
937: \be\n{4.17a}
938:   dl^2=(r_+^2+a^2)\, \left[\tilde{F}(x)\, dx^2 + \frac{d\phi^2}{\tilde{F}(x)}\right],
939: \ee
940: where
941: \be\n{4.17b}
942:   \tilde{F}(x)=\frac{1}{1-x^2} - \beta^2\, ,\hspace{1em}
943:   \beta=\frac{a}{\sqrt{r_+^2+a^2}}\, .
944: \ee
945: Here $r_+=M+\sqrt{M^2-a^2}$ gives the position of the event horizon,
946: and $M$ and $a$ are the mass and the rotation parameter of the Kerr
947: black hole. The line element (\ref{4.15},\ref{4.16}) is obtained from
948: the above by coordinate redefinition $z=\mu x$ and analytic
949: continuation $\alpha=i\beta$, with
950: $\alpha = \frac{\mu}{2\pi} (1-\frac{\mu}{\pi})^{-\frac{1}{2}}$.
951: 
952: Denote by $l_{\text{eq}}$ the proper length of the equatorial circumference,
953: and by $l_{\text{pole}}$ the proper length of a closed geodesic passing
954: through both poles $|z|=\mu$ of the black hole horizon. Then one has
955: \ba\n{4.18}
956: l_{\text{eq}}(\mu) &\approx& 2\pi\, {\sqrt{1-\mu/\pi}\over 1-\mu/(2\pi)}\, ,\\
957: l_{\text{pole}}(\mu) &\approx& 4 E\left({i\mu \over 2\pi \sqrt{1-\mu/\pi}}\right)\, ,\nonumber
958: \ea
959: where $E(k)$ is the complete elliptic integral of the second kind. One
960: has $l_{\text{eq}}(0)=l_{\text{pole}}(0)=2\pi$ and the surface is a
961: round sphere. For $\mu\to \pi$ the lengths $l_{\text{eq}}\to 0$ and
962: $l_{\text{pole}}\to \infty$.
963: 
964: \subsection{Embedding diagrams for a distorted horizon}
965: 
966: The metric (\ref{4.15}) can be obtained as an induced geometry on a
967: surface of rotation $\Sigma$ embedded in a 3 dimensional Euclidean
968: space. Let
969: \be\n{6.2}
970: dl^2=dh^2+dr^2+r^2\, d\phi^2\,
971: \ee
972: be the metric of the Euclidean space and the surface $\Sigma$ be
973: determined by an equation $h=h(r)$, then the induced metric on
974: $\Sigma$ is
975: \be\n{6.3}
976: d\sigma^2=\left[1+\left({dh\over dr} \right)^2 \right]\, dr^2+r^2\, d\phi^2\, .
977: \ee
978: By comparing this metric with (\ref{4.15}) we get
979: \be\n{6.4}
980: r={1\over \sqrt{F(z)}}\, ,
981: \ee
982: \be\n{6.5}
983: \left({dh\over dz}\right)^2+\left( {dr\over dz}\right)^2= F(z)\, .
984: \ee
985: These equations imply the following differential equation for $h(z)$
986: \be\n{6.6}
987: {dh\over dz}=\sqrt{F-{{F'}^2\over 4F^3}}\, .
988: \ee
989: 
990: Figure~\ref{embed} shows the embedding diagrams for the distorted
991: horizon surfaces of a compactified black hole for different values of
992: $\mu$. The larger is the value $\mu$ the more oblate is the surface of
993: the horizon. For large $\mu$ close to $\pi$ it has a cigar-like
994: form.
995: 
996: \subsection{$\mu\to\pi$ limit}
997: 
998: Let us now discuss the properties of the spacetime in the limiting case
999: $\mu\to\pi$. This limit can be easily taken in the series
1000: representation (\ref{3.34}) for the gravitational potential $U$. Since
1001: $\sin(\pi k)=0$ for $k>0$, only the logarithmic term survives in this
1002: limit. Thus $U(\rho,z)=\ln \rho$. Since the limiting metric is
1003: invariant under translations in $z$-direction, it has the form of the
1004: Kasner solution (\ref{4.4}) with $\mu=\pi$ and reads
1005: \be\n{6.7}
1006: ds^2=-\rho^2\, dt^2+d\rho^2+dz^2+d\phi^2\, .
1007: \ee
1008: This is a Rindler metric with two dimensions orthogonal to the
1009: acceleration direction being compactified
1010: \be\n{6.8}
1011: z\in (-\pi,\pi)\, ,\hspace{1em}\phi \in (-\pi,\pi)\, .
1012: \ee
1013: Restoring the dimensionality we can write this metric as
1014: \be\n{6.9}
1015: dS^2=-{R^2\over L^2}\, dT^2+dR^2+dZ^2+L^2 d\phi^2\, .
1016: \ee
1017: 
1018: 
1019: 
1020: \section{Discussion}\label{C}
1021: 
1022: The obtained results can be summarized as follows. If the size of a
1023: black hole is much smaller that the size of compactification, its
1024: distortion is small. The deformation which makes the horizon prolated
1025: grows with the black hole mass. For large mass $\mu \ge \pi/2$ the
1026: black hole deformation becomes profound. The pole parts of the
1027: horizon, that is parts close to $z=-\mu$ and $z=\mu$, attract one
1028: another. As a result of this attraction the Gaussian curvature of
1029: regions close to black hole poles grows, while the Gaussian curvature
1030: in the `equatorial' region falls down and the surface of the horizon
1031: is `flattened down' in this region. For large value of the mass
1032: $\mu$, the `flattening' effects occurs for a wide range of the
1033: parameter $z$. Such a black hole reminds a cigar or a part of the
1034: cylinder with two sharpened ends.
1035: 
1036: We did not include any branes in our consideration. However, we should
1037: note that the surface $Z=0$ is a solution of the Nambu-Goto action for
1038: a test brane. This can be easily seen, as the solution we discussed is
1039: symmetric around the surface $Z=0$, which implies that its extrinsic
1040: curvature vanishes there. At far distances the induced gravitational
1041: field on the $Z=0$ submanifold is asymptotically a solution of vacuum
1042: $(2+1)$-dimensional Einstein equations. It is not so for regions close
1043: to the black hole. This ``violation'' of the vacuum $(2+1)$-dimensional
1044: Einstein equations for the induced metric makes the existence of the
1045: $(2+1)$-dimensional black hole on the brane possible.
1046: 
1047: In our work we did not find any indications on instability of a black
1048: hole which might be interpreted as connected with the Gregory-Laflamme
1049: instability \cite{GrLa:93,GrLa:94}. It may not be surprising since
1050: these kind of instabilities are expected in spacetimes with higher
1051: number of dimensions (see e.g.~\cite{GrLa:88,Kol:02,Wise:02a,Wise:02b}).
1052: 
1053: \begin{figure}[t]
1054:   \centerline{\epsfig{file=stability.eps, height=2in}}
1055:   \caption{$M$ as a function of $L$ for fixed $M_0$.}
1056:   \label{LM}
1057: \end{figure}
1058: 
1059: On the other hand, a solution describing a black hole in a compactified
1060: spacetime may be unstable for a different reason. The nature of this instability is the
1061: following. In our set-up we fix a radius of compactification $L$. In a
1062: flat spacetime we can choose parameter $L$ arbitrarily and the energy
1063: of the system, being equal to zero, does not depend on this choice. The
1064: situation is different in the presence of a black hole. Consider a
1065: black hole of a given area, that is with a fixed parameter $M_0$. Since
1066: the black hole entropy, which is proportional to the area, remains
1067: unchanged for quasi-stationary adiabatic processes, one may consider
1068: different states of a black hole with a given $M_0$. $L$ plays a role
1069: of an independent parameter, specifying a solution. In particular one
1070: has
1071: \be\n{5.1}
1072:   M_0={M\, (4\pi)^{-M/(\pi L)}\over \sqrt{1-M/(\pi L)}}\, .
1073: \ee
1074: This relation shows that for fixed $M_0$ the energy of the system $M$
1075: depends on compactification radius $L$. The plot of the function $M(L)$
1076: is shown at Figure~\ref{LM}. For $L=L_*=1.345 M_0$ the mass $M$ has
1077: maximum $M=M_*=3.3877 M_0$. At the corresponding value $\mu_*=2.52$ the
1078: function $u(\mu)$ has its maximum. Thus if one starts with a system
1079: with $L>L_*$ then a positive variation of parameter $L$ will decrease
1080: the energy of the system. In this case the lowest energy state
1081: corresponds to $L\to\infty$, so that a stable solution will be an
1082: isolated Schwarzschild black hole in an empty spacetime without any
1083: compactifications. In the opposite case, $L<L_*$, the energy decreases
1084: when $L\to 0$. In this limit $M\approx \pi L$ and hence it corresponds
1085: to a limiting solution $\mu\to\pi$. The limiting metric is given by
1086: (\ref{6.9}). The corresponding spacetime is a 2-D torus
1087: compactification of the Rindler metric.
1088: 
1089: This argument, based on the energy consideration, indicates a possible
1090: instability of a compactified spacetime with a black hole with respect
1091: to compactified dimension either `unwrapping' completely or being
1092: `swallowed' by a black hole. While `unwrapping' of the extra dimension
1093: may be prevented by the usual stabilization mechanisms, the other
1094: instability regime might not be so benign. It is interesting to check
1095: whether this conjecture is correct by standard perturbation analysis.
1096: 
1097: 
1098: 
1099: \section*{Acknowledgments}
1100: This work was partly supported by the Natural Sciences and Engineering
1101: Research Council of Canada. One of the authors (V.F.) is grateful to
1102: the Killam Trust for its financial support.
1103: 
1104: 
1105: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1106: 
1107: \begin{thebibliography}{9}
1108: 
1109: \bibitem{GiWi:86}
1110: G.~W.~Gibbons and D.~L.~Wiltshire,
1111: {\it Black holes in Kaluza-Klein theory},
1112: Annals Phys.\  {\bf 167}, 201 (1986)
1113: [Erratum-ibid.\  {\bf 176}, 393 (1987)].
1114: %%CITATION = APNYA,167,201;%%
1115: 
1116: \bibitem{Lars:00}
1117: F.~Larsen,
1118: {\it Kaluza-Klein black holes in string theory},
1119: \texttt{hep-th/0002166}.
1120: %%CITATION = HEP-TH 0002166;%%
1121: 
1122: \bibitem{Myers:87}
1123: R.~C.~Myers,
1124: {\it Higher dimensional black holes in compactified space-times},
1125: Phys.\ Rev.\ D {\bf 35}, 455 (1987).
1126: %%CITATION = PHRVA,D35,455;%%
1127: 
1128: \bibitem{ChHaRe:00}
1129: A.~Chamblin, S.~W.~Hawking and H.~S.~Reall,
1130: {\it Brane-world black holes},
1131: Phys.\ Rev.\ D {\bf 61}, 065007 (2000)
1132: [\texttt{hep-th/9909205}].
1133: %%CITATION = HEP-TH 9909205;%%
1134: 
1135: \bibitem{EmHoMy:00}
1136: R.~Emparan, G.~T.~Horowitz and R.~C.~Myers,
1137: {\it Exact description of black holes on branes},
1138: JHEP {\bf 0001}, 007 (2000)
1139: [\texttt{hep-th/9911043}].
1140: %%CITATION = HEP-TH 9911043;%%
1141: 
1142: \bibitem{KuTaNa:03}
1143: H.~Kudoh, T.~Tanaka and T.~Nakamura,
1144: {\it Small localized black holes in braneworld: Formulation and numerical method},
1145: \texttt{gr-qc/0301089}.
1146: %%CITATION = GR-QC 0301089;%%
1147: 
1148: \bibitem{BW}
1149: N.~Arkani-Hamed, S.~Dimopoulos and G.~R.~Dvali,
1150: {\it The hierarchy problem and new dimensions at a millimeter},
1151: Phys.\ Lett.\ B {\bf 429}, 263 (1998)
1152: [\texttt{hep-ph/9803315}].
1153: %%CITATION = HEP-PH 9803315;%%
1154: 
1155: \bibitem{BoPe:90}
1156: A.~R.~Bogojevic and L.~Perivolaropoulos,
1157: {\it Black holes in a periodic universe},
1158: Mod.\ Phys.\ Lett.\ A {\bf 6}, 369 (1991).
1159: %%CITATION = MPLAE,A6,369;%%
1160: 
1161: \bibitem{HaOb:02}
1162: T.~Harmark and N.~A.~Obers,
1163: {\it Black holes on cylinders},
1164: JHEP {\bf 0205}, 032 (2002)
1165: [\texttt{hep-th/0204047}].
1166: %%CITATION = HEP-TH 0204047;%%
1167: 
1168: \bibitem{IsKh:64}
1169: W.~Israel and K.~A.~Khan,
1170: {\it Collinear particles and Bondi dipoles in General Relativity},
1171: Nouvo Cimento {\bf 33}, 331 (1964).
1172: 
1173: \bibitem{KoNi:94}
1174: D.~Korotkin and H.~Nicolai,
1175: {\it A periodic analog of the Schwarzschild solution},
1176: \texttt{gr-qc/9403029}.
1177: %%CITATION = GR-QC 9403029;%%
1178: 
1179: \bibitem{GeHa:81}
1180: R.~Geroch and J.~B.~Hartle,
1181: {\it Distorted black holes},
1182: J.\ Math.\ Phys.\ {\bf 23} 680 (1981).
1183: 
1184: \bibitem{FaKr:01}
1185: S.~Fairhurst and B.~Krishnan,
1186: {\it Distorted black holes with charge},
1187: Int.\ J.\ Mod.\ Phys.\ D {\bf 10}, 691 (2001)
1188: [\texttt{gr-qc/0010088}].
1189: %%CITATION = GR-QC 0010088;%%
1190: 
1191: \bibitem{Yaza:01}
1192: S.~S.~Yazadjiev,
1193: {\it Distorted charged dilaton black holes},
1194: Class.\ Quant.\ Grav.\  {\bf 18}, 2105 (2001)
1195: [\texttt{gr-qc/0012009}].
1196: %%CITATION = GR-QC 0012009;%%
1197: 
1198: \bibitem{EmRe:02}
1199: R.~Emparan and H.~S.~Reall,
1200: {\it Generalized Weyl solutions},
1201: Phys.\ Rev.\ D {\bf 65}, 084025 (2002)
1202: [\texttt{hep-th/0110258}].
1203: %%CITATION = HEP-TH 0110258;%%
1204: 
1205: \bibitem{ShSh:00}
1206: T.~Shiromizu and M.~Shibata,
1207: {\it Black holes in the brane world: Time symmetric initial data},
1208: Phys.\ Rev.\ D {\bf 62}, 127502 (2000)
1209: [\texttt{hep-th/0007203}].
1210: %%CITATION = HEP-TH 0007203;%%
1211: 
1212: \bibitem{SoPi:02}
1213: E.~Sorkin and T.~Piran,
1214: {\it Initial data for black holes and black strings in 5d},
1215: \texttt{hep-th/0211210}.
1216: %%CITATION = HEP-TH 0211210;%%
1217: 
1218: \bibitem{FrNo:98}
1219: V.~P.~Frolov and I.~D.~Novikov,
1220: {\it Black hole physics: Basic concepts and new developments},
1221: Kluwer Academic Publ.\ (1998).
1222: 
1223: \bibitem{Kasner}
1224: E.~Kasner,
1225: {\it Geometrical theorems on Einstein's cosmological equations},
1226: Am.\ J.\ Math.\ {\bf 43} 217 (1921).
1227: 
1228: \bibitem{PBM}
1229: A.~P.~Prudnikov, Yu.~A.~Brychkov, and O.~I.~Marichev,
1230: {\it Integrals and series} v.I,
1231: Gordon and Breach Science Publishers (1986).
1232: 
1233: \bibitem{GrLa:88}
1234: R.~Gregory and R.~Laflamme,
1235: {\it Hypercylindrical black holes},
1236: Phys.\ Rev.\ D {\bf 37}, 305 (1988).
1237: %%CITATION = PHRVA,D37,305;%%
1238: 
1239: \bibitem{GrLa:93}
1240: R.~Gregory and R.~Laflamme,
1241: {\it Black strings and p-branes are unstable},
1242: Phys.\ Rev.\ Lett.\  {\bf 70}, 2837 (1993)
1243: [\texttt{hep-th/9301052}].
1244: %%CITATION = HEP-TH 9301052;%%
1245: 
1246: \bibitem{GrLa:94}
1247: R.~Gregory and R.~Laflamme,
1248: {\it The Instability of charged black strings and p-branes},
1249: Nucl.\ Phys.\ B {\bf 428}, 399 (1994)
1250: [\texttt{hep-th/9404071}].
1251: %%CITATION = HEP-TH 9404071;%%
1252: 
1253: \bibitem{Kol:02}
1254: B.~Kol,
1255: {\it Topology change in general relativity and the black-hole black-string transition},
1256: \texttt{hep-th/0206220}.
1257: %%CITATION = HEP-TH 0206220;%%
1258: 
1259: \bibitem{Wise:02a}
1260: T.~Wiseman,
1261: {\it From black strings to black holes},
1262: Class.\ Quant.\ Grav.\  {\bf 20}, 1177 (2003)
1263: [\texttt{hep-th/0211028}].
1264: %%CITATION = HEP-TH 0211028;%%
1265: 
1266: \bibitem{Wise:02b}
1267: T.~Wiseman,
1268: {\it Static axisymmetric vacuum solutions and non-uniform black strings},
1269: Class.\ Quant.\ Grav.\  {\bf 20}, 1137 (2003)
1270: [\texttt{hep-th/0209051}].
1271: %%CITATION = HEP-TH 0209051;%%
1272: 
1273: \end{thebibliography}
1274: 
1275: 
1276: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1277: 
1278: 
1279: \end{document}
1280: 
1281: 
1282: 
1283: 
1284: 
1285: