1: %\documentclass[twocolumn,showpacs,preprintnumbers,amssymb,nofootinbib]{revtex4}
2: %\documentclass
3: %[preprint,showpacs,preprintnumbers,amsmath,amssymb,nofootinbib]{revtex4}
4: \documentclass[showpacs,preprintnumbers,amsmath,amssymb,nofootinbib]{revtex4}
5:
6: \usepackage{graphicx,epsf, epsfig, psfig, amssymb}% Include figure files
7: \usepackage{bm}
8: % Include bold math: \bm{} creates bold letters and symbols in math mode
9:
10: \def\h{\hfill\break\noindent}
11: \def\nn{\nonumber}
12: \def\ver{\vskip 12pt}
13: \def\lll{\left[}
14: \def\rrr{\right]}
15: \def\bl{\Biggl[}
16: \def\br{\Biggr]}
17: \def\gbl{\Biggl\{}
18: \def\gbr{\Biggr\}}
19:
20: \def\pps{\Psi_{\ell m}}
21: \def\pa{\partial}
22: \def\Y{Y_{\ell m}(\vartheta,\varphi)}
23: \def\YS{Y^*_{\ell m}(\vartheta,\varphi)}
24: \def\Yt{Y_{\ell m,\vartheta }(\vartheta,\varphi)}
25: \def\Ytt{Y_{\ell m,\vartheta \vartheta}(\vartheta,\varphi)}
26: \def\YE{Y_{\ell m}}
27: \def\YSE{Y^*_{\ell m}}
28: \def\YtE{Y_{\ell m,\vartheta }}
29: \def\YttE{Y_{\ell m,\vartheta \vartheta}}
30: \def\EM{e^{-\ii\omega t}}
31: \def\EP{e^{+\ii\omega t}}
32: \def\N{N_{\ell m}(t,r)}
33: \def\T{T_{\ell m}(t,r)}
34: \def\V{V_{\ell m}(t,r)}
35: \def\L{L_{\ell m}(t,r)}
36: \def\hh{h^{0}_{\ell m}(t,r)}
37: \def\hs{h^{1}_{\ell m}(t,r)}
38: \def\msun{M_\odot}
39: \def\httw{h^{TT}_{\vartheta \vartheta}(\omega,r,\vartheta,\varphi)}
40: \def\htpw{h^{TT}_{\vartheta \varphi}(\omega,r,\vartheta,\varphi)}
41: \def\htt{h^{TT}_{\vartheta \vartheta}(t,r,\vartheta,\varphi)}
42: \def\htp{h^{TT}_{\vartheta \varphi}(t,r,\vartheta,\varphi)}
43: \def\be{\begin{equation}}
44: \def\ee{\end{equation}}
45: \def\beq{\begin{eqnarray}}
46: \def\eeq{\end{eqnarray}}
47: \def\lm{{\ell m}}
48: \def\ii{{\rm i}}
49: \def\o{\omega}
50: \def\IL{\relax{\rm I\kern-.18em L}}
51: \def\tT{{\tilde T}}
52: \def\nn{\nonumber}
53: \def\f{\frac}
54:
55: \begin{document}
56:
57: \title{Quasinormal modes and classical wave propagation in
58: analogue black holes}
59: %\title{Quasinormal modes and classical wave propagation in
60: %analogue acoustic black holes}
61:
62: \author{Emanuele Berti}
63: \email{berti@iap.fr} \affiliation{McDonnell Center for the Space Sciences,
64: Department of Physics, Washington University, St. Louis, Missouri 63130, USA}
65:
66: \author{Vitor Cardoso}
67: \email{vcardoso@teor.fis.uc.pt} \affiliation{Centro de F\'{\i}sica
68: Computacional, Universidade de Coimbra, P-3004-516 Coimbra,
69: Portugal}
70:
71: \author{Jos\'e P. S. Lemos}
72: \email{lemos@fisica.ist.utl.pt} \affiliation{Centro
73: Multidisciplinar de Astrof\'{\i}sica - CENTRA, Departamento de
74: F\'{\i}sica, Instituto Superior T\'ecnico, Av. Rovisco Pais 1,
75: 1049-001 Lisboa, Portugal}
76:
77: \date{\today}
78:
79: \begin{abstract}
80:
81: Many properties of black holes can be studied using acoustic analogues
82: in the laboratory through the propagation of sound waves. We
83: investigate in detail sound wave propagation in a rotating acoustic
84: $(2+1)$-dimensional black hole, which corresponds to the ``draining
85: bathtub'' fluid flow. We compute the quasinormal mode frequencies of
86: this system and discuss late-time power-law tails. Due to the presence
87: of an ergoregion, waves in a rotating acoustic black hole can be
88: superradiantly amplified. We compute superradiant reflection coefficients
89: and instability timescales for the acoustic black hole bomb, the
90: equivalent of the Press-Teukolsky black hole bomb. Finally we discuss
91: quasinormal modes and late-time tails in a non-rotating canonical
92: acoustic black hole, corresponding to an incompressible, spherically
93: symmetric $(3+1)$-dimensional fluid flow.
94:
95: \end{abstract}
96:
97: \pacs{04.70.-s, 43.20.+g, 04.80.Cc}
98:
99: \maketitle
100:
101: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
102: \section{Introduction}
103: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
104:
105: Black holes are one of the most fascinating predictions of general
106: relativity. Being a solution of the classical Einstein field equations
107: in vacuum, they are the simplest objects which can be built out of
108: spacetime itself. Hawking \cite{hawking} showed that when quantum
109: effects are taken into account, black holes are not really black: they
110: slowly evaporate by emitting an almost thermal radiation. Hawking's
111: prediction has been tested time and again in very different ways. It
112: is now clear that the appearance of Hawking radiation does not depend
113: on the dynamics of the Einstein equations, but only on their
114: kinematical structure, and more specifically on the existence of an
115: apparent horizon \cite{visserlaws,visserhawking}. The discovery of
116: Hawking radiation uncovered a number of fundamental questions: among
117: them the information puzzle, the issue of the black hole final state,
118: and so on. For these reasons an experimental verification of the
119: Hawking effect would be of the utmost importance. Unfortunately
120: astrophysical black holes, having a temperature much smaller than the
121: temperature of the cosmic microwave background, accrete matter more
122: efficiently than they evaporate. An alternative could be provided by
123: the existence of large extra dimensions: if gravity is effective at
124: the TeV scale, black holes could be produced in particle accelerators
125: \cite{giddings}. However, even if these mini-black holes are actually
126: produced in colliders, it is unlikely that they will yield any
127: conclusive evidence of the existence of Hawking radiation.
128:
129: Prospects for detecting Hawking radiation changed when Unruh
130: \cite{unruh} realized that the basic ingredients of Hawking radiation
131: can experimentally be reproduced in the laboratory. Since Hawking
132: radiation crucially depends on the existence of an apparent horizon,
133: the experimental setup should display the essential features that
134: define apparent horizons in general relativity. Unruh considered
135: precisely such a system: a fluid moving with a space-dependent
136: velocity, for example water flowing through a nozzle. Where the fluid
137: velocity exceeds the sound velocity we get the equivalent of an
138: apparent horizon for sound waves. This is the acoustic analogue of a
139: black hole, or a {\it dumb hole}. Following on Unruh's dumb hole
140: proposal many different kinds of analogue black holes have been
141: devised, based on condensed matter physics, slow light etcetera
142: \cite{novello,visser}. At present the Hawking temperatures associated
143: to these analogues are too low to be detectable, but the situation is
144: likely to change in the near future \cite{barcelo,unruh2}.
145:
146: In order to detect Hawking radiation, a full understanding of the {\it
147: classical} physics of analogue black holes is necessary. Not only must
148: one control what happens in the experimental situation, but the
149: understanding of classical phenomena may bring clues on how to favor
150: the probabilities to detect Hawking radiation. It is also worth
151: stressing that some purely classical phenomena shed light on quantum
152: aspects of (analogue and general-relativistic) black hole physics. For
153: example, positive and negative norm mixing at the horizon leads to
154: non-trivial Bogoliubov coefficients in the calculations of Hawking
155: radiation \cite{corleyjacobson}; superradiant instabilities of the
156: Kerr metric are related to the quantum process of Schwinger pair
157: production \cite{schwinger}; and more speculatively (classical) highly
158: damped black hole oscillations could be related to area quantization
159: \cite{hod}.
160:
161: In this paper we carry out a comprehensive study of wave propagation
162: in analogue black holes. We consider in detail two acoustic black hole
163: metrics: the ``draining bathtub'' model for a $(2+1)$-dimensional
164: fluid flow and the ``canonical'' metric for a non-rotating,
165: spherically symmetric $(3+1)$-dimensional acoustic black hole
166: \cite{visser}. For each acoustic metric we compute the characteristic
167: oscillation frequencies using a WKB approach. These complex
168: characteristic frequencies, called quasinormal modes (QNMs), play a
169: very important role in the classical physics of black holes. They
170: govern the late time behavior of waves propagating outside the black
171: hole: any perturbation of the black hole, after an initial transient,
172: will damp exponentially in the so-called ``ringdown phase''. The
173: frequencies and damping times of this ringdown signal depend only on
174: the black hole parameters, such as the mass, charge and angular
175: momentum \cite{kokkotas}, and can therefore be used to estimate black
176: hole parameters from observational data \cite{echeverria}.
177: Therefore, in cases where a ``formal'' definition of (say) the mass
178: and angular momentum of acoustic black holes is missing, quasinormal
179: (QN) frequencies may be used to define these quantities {\it
180: operationally}. According to recent speculations, highly damped QNMs
181: may yield some information on quantum properties of a black hole. In
182: particular, it has been conjectured that highly damped QN frequencies
183: could be linked to black hole area quantization \cite{hod}. Motivated
184: by these conjectures we also study highly damped QNMs of acoustic
185: black holes. We find that for a $(2+1)$-dimensional acoustic black
186: hole there are no asymptotic QN frequencies, whereas for the
187: $(3+1)$-dimensional canonical black hole QN frequencies are
188: asymptotically given by $4\pi \omega =\log{[(3-\sqrt{5})/2]}
189: -\ii(2n+1)\pi$.
190: This result does not seem to support any of the recent
191: conjectures, but perhaps this is no surprise, since Hod's argument
192: relies heavily on black hole thermodynamics. Even the very formulation
193: of the laws of black hole thermodynamics for analogue black holes is a
194: non-trivial matter \cite{visserlaws}.
195:
196: After the exponential decay characteristic of the ringdown phase,
197: black hole perturbations decay with a power-law tail \cite{price1} due
198: to backscattering off the background curvature. Here we compute the
199: late-time tails of wave propagation in the draining bathtub and
200: canonical acoustic black hole metrics. We show that for the
201: $(2+1)$-dimensional fluid flow the field falloff at very late times is
202: of the form $\Psi\sim t^{-(2m+1)}$, where $m$ is the angular quantum
203: number. This time exponent is characteristic of any
204: $(2+1)$-dimensional flow, and not just of a black hole. For the
205: $(3+1)$-dimensional canonical acoustic geometry, the field falloff is
206: of the form $\Psi \sim t^{-(2l+6)}$.
207:
208: The rotating draining bathtub metric, possessing an ergoregion, can
209: display the phenomenon of superradiance
210: \cite{schutzhold,basak1,basak2}. We compute reflection coefficients
211: for this superradiant scattering by numerical integration of the
212: relevant equations. Enclosing the acoustic black hole by a reflecting
213: mirror we can exploit superradiance to destabilize the system, making
214: an initial perturbation grow exponentially with time: we have an
215: ``acoustic black hole bomb'' \cite{pressteu,bhb,putten}. We compute
216: analytically and numerically the frequencies and growing timescales
217: for this instability. An interesting feature of acoustic geometries
218: is that the acoustic black hole {\it spin} can be varied independently
219: of the black hole {\it mass}. Therefore, at variance with the Kerr
220: metric, the spin can be made (at least in principle) very large, and
221: rotational superradiance in acoustic black holes can be very
222: efficient. It should not be difficult to set up an experimental
223: apparatus to observe an acoustic black hole bomb in the lab. A
224: concrete example of an experimental setup in which this idea can be
225: realized with moderate experimental effort has been fully worked out
226: by Sch\"utzhold and Unruh \cite{schutzhold}, and is based on the study
227: of gravity waves in shallow water.
228:
229: The paper is organized as follows. In Section \ref{bathtub} we discuss
230: the $(2+1)$-dimensional draining bathtub metric. We first introduce
231: the formalism describing sound propagation in this acoustic metric,
232: and shortly describe the possible experimental setup (gravity waves in
233: a shallow basin). Then we compute QNMs, discuss late-time tails,
234: introduce the phenomenon of superradiant amplification and quantify
235: the timescales for the acoustic black hole bomb instability. In
236: Section \ref{canonical} we repeat the analysis for the canonical
237: $(3+1)$-dimensional acoustic black hole metric (of course, the absence
238: of an ergoregion means that we have no superradiance in this
239: case). The conclusions follow in Section \ref{conclusions}.
240:
241: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
242: \section{Draining bathtub: a rotating acoustic black hole}\label{bathtub}
243: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
244:
245: In this Section we consider a simple ``draining bathtub'' model, first
246: introduced in \cite{visser}, for a rotating acoustic black hole. We
247: write down the acoustic metric and the wave equation describing sound
248: propagation in this model, and describe a concrete example for a
249: possible experimental setup. We compute QNMs and discuss the late-time
250: tail behavior of the system. Due to the presence of an ergoregion,
251: sound waves in this acoustic black hole can be superradiantly
252: amplified. We quantify this amplification and discuss the possibility
253: to build an acoustic black hole bomb in the lab.
254:
255: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
256: \subsection{Formalism and basic equations}
257: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
258:
259: Consider a fluid having (background) density $\rho$. Assume the fluid
260: to be locally irrotational (vorticity free), barotropic and
261: inviscid. From the equation of continuity, the radial component of the
262: fluid velocity satisfies $\rho v^r\sim 1/r$. Irrotationality implies
263: that the tangential component of the velocity satisfies $v^\theta\sim
264: 1/r$. By conservation of angular momentum we have $\rho v^\theta\sim
265: 1/r$, so that the background density of the fluid $\rho$ is
266: constant. In turn, this means that the background pressure $p$ and the
267: speed of sound $c$ are constants. The acoustic metric describing the
268: propagation of sound waves in this ``draining bathtub'' fluid flow is
269: \cite{visser}:
270: %
271: \begin{equation}
272: ds^2=
273: -\left (c^2-\frac{A^2+B^2}{r^2} \right )dt^2+\frac{2A}{r}drdt-2Bd\phi
274: dt+dr^2+
275: r^2d\phi^2.
276: \label{metric1}
277: \end{equation}
278: Here $A$ and $B$ are arbitrary real positive constants related to the
279: radial and angular components of the background fluid velocity:
280: %
281: \be
282: {\vec v}=\frac{-A {\hat r}+B {\hat \theta}}{r}\,.
283: \ee
284: %
285: This flow velocity can be obtained as the gradient of a velocity
286: potential, $\vec v=\nabla \psi$, where
287: %
288: \be\label{vpot}
289: \psi=-A\log (r/a)+B\phi\,.
290: \ee
291: %
292: and $a$ is some (irrelevant) length scale.
293:
294: In the non-rotating limit $B=0$ the metric (\ref{metric1}) reduces to
295: a standard Painlev\'e-Gullstrand-Lema\^itre type metric
296: \cite{PGL}. The acoustic event horizon is located at $r_H=A/c$, and
297: the ergosphere forms at $r_{ES}=(A^2+B^2)^{1/2}/c$.
298: %
299: Unruh \cite{unruh} first realized that the propagation of a sound wave
300: in a barotropic inviscid fluid with irrotational flow is described by
301: the Klein-Gordon equation $\nabla_{\mu}\nabla^{\mu}\Psi=0$ for a
302: massless field $\Psi$ in a Lorentzian acoustic geometry, which in our
303: case takes the form (\ref{metric1}). In our acoustic geometry we can
304: separate variables by the substitution
305: %
306: \be
307: \Psi(t,r,\phi)=R(r)e^{\ii (m\phi-\omega t)}\,.
308: \ee
309: %
310: Then we obtain a simple ordinary differential equation for the radial
311: variable:
312: %
313: \be
314: R_{,rr}+P_1(r)R_{,r}+Q_1(r)R=0\,,
315: \ee
316: where
317: \beq
318: P_1(r)&=&\f{A^2+r^2c^2+2\ii A(Bm-r^2\omega)}{r(r^2c^2-A^2)}\,,\nn\\
319: Q_1(r)&=&-\f{2\ii A Bm-B^2m^2+c^2m^2r^2+2Bm\omega r^2-r^4\omega^2}
320: {r^2(r^2c^2-A^2)}\,.
321: \eeq
322: %
323: Notice that if we take the incompressible fluid limit $c\rightarrow
324: \infty$ we get well known equations from fluid dynamics
325: \cite{chandra,drazin}. We now introduce a tortoise coordinate $r_*$
326: defined by the condition
327: %
328: \be
329: \frac{dr_*}{dr}=\Delta\,,
330: \ee
331: where $\Delta\equiv (1-A^2/c^2r^2)^{-1}$. Explicitly,
332: \be
333: r_*=r+\f{A}{2c}\log\left|\f{cr-A}{cr+A}\right|\,.
334: \ee
335: Setting $R=ZH$ we get:
336: \be\label{ZH}
337: Z\Delta^2 H_{,r_* r_*}
338: +\left[\Delta(2Z_{,r}+P_1Z)+\Delta'Z\right]H_{,r_*}+
339: \left(Z_{,rr}+P_1Z_{,r}+Q_1Z\right)H=0\,.
340: \ee
341: %
342: To obtain a Schr\"odinger-like equation we impose the coefficient of
343: $H_{,r_*}$ to be zero:
344: %
345: \be
346: Z_{,r}+\f{1}{2}\left[
347: \f{(c^2r^2-A^2)+2\ii A(Bm-r^2\omega)}{r(c^2r^2-A^2)}
348: \right]Z=0\,.
349: \ee
350: %Set $Z=r^{1/2}e^{g(r)}$; then
351: %\be
352: %g_{,r}+\f{(c^2r^2-A^2)+\ii A(Bm-r^2\omega)}{r(c^2r^2-A^2)}=0
353: %\ee
354: %Making the ansatz $g=k\ln r+h \ln (c^2r^2-A^2)$ we get
355: %\be
356: %k=\f{\ii Bm}{A}-1
357: %, \qquad h=\f{\ii(A^2\omega-Bmc^2)}{2Ac^2}.
358: %\ee
359: A solution of this equation can easily be found:
360: \be
361: Z=r^{1/2}\exp{\left[
362: \left(\f{\ii Bm}{A}-1\right)\ln r
363: +\f{\ii(A^2\omega-Bmc^2)}{2Ac^2}\ln (c^2r^2-A^2)
364: \right]}\,.
365: \ee
366: Replacing this solution in Eq. (\ref{ZH}) we finally get the wave equation:
367: \be
368: H_{,r_* r_*}+
369: \left\{\f{1}{c^2}\left(\omega-\f{Bm}{r^2}\right)^2-
370: \left(\f{c^2r^2-A^2}{c^2r^2}\right)
371: \left[\f{1}{r^2}\left(m^2-\f{1}{4}\right)+\f{5A^2}{4r^4 c^2}\right]\right\}
372: H=0\,.
373: \label{waveequation}
374: \ee
375:
376: The derivation given here is very similar to the one in \cite{basak1},
377: but we have corrected some typos in that paper. In particular notice
378: that the potential is the same as equation (10) in \cite{basak1},
379: except for a factor $1/4$ in the last term. Some physical properties
380: of our ``draining bathtub'' metric are more apparent if we cast the
381: metric in a Kerr-like form performing the following coordinate
382: transformation (where again we correct some typos in \cite{basak2}):
383: \begin{equation}
384: dt\rightarrow d\tilde{t}+\frac{Ar}{r^2c^2-A^2}dr\,,\qquad
385: d\phi\rightarrow d\tilde{\phi}+\frac{BA}{r(r^2c^2-A^2)}dr\,.
386: \label{coordtransf}
387: \end{equation}
388: Then the effective metric takes the form
389: \begin{equation}
390: ds^2=
391: -\left(1-\frac{A^2+B^2}{c^2r^2} \right)c^2 d\tilde{t}^2+
392: \left(1-\frac{A^2}{c^2r^2} \right )^{-1}dr^2
393: -2B d\tilde{\phi}d\tilde{t}+r^2d\tilde{\phi}^2\,.
394: \label{metric2}
395: \end{equation}
396: %
397: Notice an important difference between this acoustic metric and the
398: Kerr metric: in the ($t,t$) component of the metric (\ref{metric2})
399: the parameters $A$ and $B$ appear as a {\it sum} of squares. This
400: means that, {\it at least in principle}, there is no upper bound for
401: the rotational parameter $B$ in the acoustic black hole metric,
402: contrary to what happens in the Kerr geometry. Separating variables by
403: the substitution
404: %
405: \be
406: \Psi(\tilde{t},r,\tilde{\phi})=\sqrt{r} H(r)e^{\ii (m\tilde{\phi}-\omega \tilde{t})}\,,
407: \ee
408: %
409: one can show that the radial function $H(r)$ is again a solution of
410: equation (\ref{waveequation}).
411:
412: The wave equation (\ref{waveequation}) can be recast in a more
413: convenient form by the following rescaling: $\hat r=rA/c$,
414: $\hat{\omega}=\omega A/c^2$, $\hat{B}=B/A$. We get
415: %
416: \be
417: H_{,\hat r_* \hat r_*}+Q H=0\,,
418: \label{waveequation2}
419: \ee
420: %
421: where the generalized potential
422: \be\label{Qdef}
423: Q\equiv
424: \left\{ \left(\hat{\omega}-\f{\hat{B}m}{\hat r^2}\right)^2-V \right\}\,,
425: \qquad
426: V\equiv \left(\f{\hat r^2-1}{\hat r^2}\right)
427: \left[\f{1}{\hat r^2}\left(m^2-\f{1}{4}\right)+\f{5}{4\hat r^4}\right]\,.
428: \ee
429:
430: The rescaling effectively sets $A=c=1$ in the original wave equation,
431: and picks units such that the acoustic horizon $\hat r_H=1$. From now
432: on we shall omit hats in all quantities (unless otherwise stated). The
433: rescaled wave equation (\ref{waveequation2}) will be the starting
434: point of our analysis of QNMs, late-time tails and superradiant
435: phenomena. Before giving details of this analysis we describe a
436: possible experimental setup in which all of the classical physics we
437: are going to discuss could, at least in principle, be reproduced.
438:
439: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
440: \subsection{A possible experimental setup}\label{gwanal}
441: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
442:
443: The acoustic metric (\ref{metric1}) can be realized in a simple
444: experimental setup, that was described in detail by Sch\"utzhold and
445: Unruh \cite{schutzhold}. In this setup it no longer describes a sonic
446: analogue but rather a shallow basin gravity wave analogue of a black
447: hole. The idea is to use gravity waves in a viscosity free,
448: incompressible liquid with irrotational flow: under appropriate
449: circumstances, one can envisage the use of common fluids like water or
450: mercury. Sch\"utzhold and Unruh assumed a shallow water, long
451: wavelength approximation: the gravity wave amplitude $\delta h$, their
452: wavelength $\lambda$ and the depth of the basin $h_B$ are such that
453: $\delta h\ll h_B\ll \lambda$. Relaxing the assumptions that the bottom
454: of the tank and the background flow surfaces are flat and parallel,
455: they showed that the most general rotationally symmetric and locally
456: irrotational background flow profile can be described precisely by the
457: draining bathtub metric (\ref{metric1}) when the radial slope of the
458: bottom of the tank -- that in cylindrical coordinates $(z,~r,~\phi)$
459: will be described by some function $f(r)$ -- is small: $f'(r)\ll
460: 1$. In this gravity wave black hole analogue the constants $A$ and $B$
461: are proportional to the radial and tangential components of the
462: background flow velocity:
463: %
464: \be
465: v^\phi=\frac{B}{r^2}\,,
466: \qquad
467: v^r=-\frac{A h_\infty}{rh\sqrt{1+f'(r)^2}}\,,
468: \ee
469: %
470: where $h_\infty$ is the height of the tank far from the black hole,
471: and the slope of the tank satisfies the relation
472: %
473: \be
474: f(r)=-\frac{(A^2+B^2)}{gr^2}\,.
475: \ee
476: %
477: In the previous equations $g$ is the gravitational acceleration,
478: related to the constant $c$ of the acoustic black hole metric
479: (\ref{metric1}) by the relation
480: %
481: \be
482: c=\sqrt{gh_\infty}\,.
483: \ee
484: %
485: One of the main advantages of this acoustic black hole is apparent
486: from this equation: the speed of the gravity waves can simply be tuned
487: to one's needs by adjusting the height of the basin
488: $h_\infty$. Another advantage, that we will not exploit here, is that
489: the inclusion of surface tension and viscosity can be used to
490: manipulate the waves' dispersion relation. To be concrete we will
491: sometimes consider the following plausible choice of physical
492: parameters, as suggested in \cite{schutzhold}: gravity waves of
493: amplitude $\delta h\sim 1$ mm and wavelength $\lambda\sim 10$ cm; a
494: tank of height $h_\infty\sim 1$ cm (so that $c\sim 0.31$ m~s$^{-1}$);
495: and a typical characteristic size of the acoustic black hole horizon
496: $r_H\sim 1$ m, corresponding to $A=c r_H\sim 0.31$ m$^2$~s$^{-1}$.
497:
498: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
499: \subsection{Quasinormal modes}\label{dbqnms}
500: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
501:
502: Numerical and analytical studies of a fairly general class of initial
503: data show that the evolution of the perturbations of a black hole
504: spacetime can roughly be divided in three parts: (i) The first part is
505: the prompt response at very early times. In this phase, which is the
506: obvious counterpart of light cone propagation, the form of the signal
507: depends strongly on the initial conditions. (ii) At intermediate times
508: the signal is dominated by an exponential decay, whose frequencies and
509: damping times are determined by the black hole QNMs. In this
510: ``ringdown'' phase the signal depends entirely on the black hole
511: parameters (typically mass, charge and angular momentum). (iii) Due to
512: backscattering off the spacetime curvature, at late times the
513: propagating wave leaves a ``tail'' behind, usually a power law falloff
514: of the field. This power law seems to be highly independent of the
515: initial data, and persists even if there is no black hole horizon. In
516: fact it only depends on the asymptotic far region. In the following we
517: study the QNM ringing phase of our acoustic black hole metric; then we
518: will consider its late-time behavior.
519:
520: The characteristic QNMs of the rotating acoustic black hole can be
521: defined in the usual way, imposing appropriate boundary conditions and
522: solving the corresponding eigenvalue problem. Close to the event
523: horizon the solutions of equation (\ref{waveequation2}) behave as
524: %
525: \begin{equation}
526: H \sim e^{\pm \ii\left (\omega-Bm\right)r_* }\,.
527: \label{bound1}
528: \end{equation}
529: %
530: Classically, only ingoing waves -- that is, waves falling into the
531: black hole -- should be present at the horizon. This means (according
532: to our conventions on the time dependence of the perturbations) that
533: we must choose the minus sign in the exponential. At spatial
534: infinity the solutions of (\ref{waveequation2}) behave as
535: %
536: \begin{equation}
537: H \sim e^{\pm \ii\omega r_*}\,.
538: \label{bound2}
539: \end{equation}
540: %
541: In this case we require that only outgoing waves (waves leaving the
542: domain under study) should be present, and correspondingly choose the
543: plus sign in the exponential. This boundary condition at infinity may
544: be cause for objections. Indeed, no actual physical apparatus will be
545: accurately described by these boundary conditions: a real acoustic
546: black hole experiment will certainly not extend out to
547: infinity. However, we may imagine using some absorbing device to
548: simulate the ``purely outgoing'' wave conditions at infinity (for
549: another example in which an absorbing device modeling spatial infinity
550: could be required, cf. Section XI of \cite{schutzhold} -- in
551: particular their Fig. 5). In any event, later on we shall consider the
552: alternative possibility of ``boxed'' boundary conditions, describing a
553: closed system. Boxed boundary conditions were also considered in
554: \cite{bhb}.
555:
556: For assigned values of the rotational parameter $B$ and of the angular
557: index $m$ there is a discrete (and infinite) set of QN frequencies,
558: $\omega_{QN}$, satisfying the wave equation (\ref{waveequation2}) with
559: boundary conditions specified by Eqs. (\ref{bound1}) and
560: (\ref{bound2}). The QN frequencies are in general complex numbers, the
561: imaginary part describing the decay or growth of the perturbation,
562: because the time dependence is given by $e^{-\ii \omega t}$. We expect
563: the black hole to be stable against small perturbations, and therefore
564: $\omega_{QN}$ is expected to have a negative imaginary part, so that
565: the perturbation decays exponentially as time goes by. We could not
566: prove stability of the draining bathtub metric in general, but we
567: managed to derive upper bounds on the frequencies of unstable modes
568: (if they exist at all). We give the derivation of these upper bounds
569: in Appendix \ref{unstable}. As usual, we will order the QN frequencies
570: $\omega_{QN}$ according to the absolute value of their imaginary part:
571: the fundamental mode (labeled by an integer $n=0$) will have the
572: smallest imaginary part (in modulus), and so on.
573:
574: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
575: \subsubsection{Slowly decaying modes of non-rotating black holes}
576: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
577:
578: The lowest QNMs control the ringing behavior of any classical
579: perturbation outside the black hole. Higher overtones, having a larger
580: imaginary part, are damped more quickly and play a negligible role. In
581: particular, it is known \cite{kokkotas} that the fundamental ($n=0$)
582: QNM effectively determines the response of the black hole to exterior
583: perturbations. For these slowly damped QNMs, a WKB approximation -- as
584: developed by Schutz, Will and others \cite{willwkb,will,kokkotaswkb}
585: -- is accurate enough. As a convergence check, we will sometimes
586: extend the WKB treatment to sixth order \cite{konoplyawkb}. The WKB
587: method demands that the generalized potential $Q$ defined by Eq.
588: (\ref{Qdef}) has a single maximum outside the horizon. For
589: non-rotating black holes ($B=0$) this is certainly true provided $m$
590: is not zero, and one can easily extract the lowest QN frequencies in
591: this case. We show them in Table \ref{tab:fundnonrot}, where we also
592: show the convergence of the WKB scheme as the order of approximation
593: increases. For $m=0$ the situation is not so simple, and we haven't
594: been able to extract the QN frequencies using this method, because the
595: potential possesses two extrema. Due to the symmetry of the potential
596: (\ref{Qdef}), QN frequencies of non-rotating black holes are
597: independent of the sign of $m$ (to any positive QN frequency for
598: positive $m$ corresponds a negative QN frequency for negative $m$).
599:
600: \begin{table}
601: \centering
602: \caption{\label{tab:fundnonrot} The fundamental ($n=0$) QN frequencies
603: for the non-rotating acoustic black hole, using three WKB
604: computational schemes. $\omega _{QN}^{(1)}$ is the result for the QN
605: frequency using only the lowest approximation \cite{willwkb}, $\omega
606: _{QN}^{(3)}$ is the value obtained using 3rd order improvements
607: \cite{will}, and finally $\omega _{QN}^{(6)}$ was computed using 6th
608: order corrections \cite{konoplyawkb}. Notice how for $m>1 $ the three
609: schemes yield very similar answers.}
610: \vskip 12pt
611: \begin{tabular}{@{}c|c|c|c@{}}
612: \hline
613: \hline
614: $m$ &$\omega _{QN}^{(1)}$ &$\omega _{QN}^{(3)}$ &$\omega _{QN}^{(6)}$\\
615: \hline
616: \hline
617: %0 & & 0.035-0.609i & \\
618: 1 & 0.696-0.353i & 0.321-0.389i &0.427-0.330i\\
619: 2 & 1.105-0.349i & 0.940-0.353i &0.945-0.344i\\
620: 3 & 1.571-0.351i & 1.465-0.353i &1.468-0.352i \\
621: 4 & 2.054-0.352i & 1.975-0.353i &1.976-0.353i \\
622: \hline
623: \hline
624: \end{tabular}
625: \end{table}
626: %
627: From Table \ref{tab:fundnonrot} we see that the imaginary part is
628: nearly constant as a function of $m$, whereas the real part
629: approximately scales with $m$. Indeed, in the limit of large $m$ one
630: can show directly from the WKB formula \cite{willwkb} that
631: $\omega_{QN}$ behaves as
632: \begin{equation}
633: \omega \sim \frac{m}{2}-i\frac{2n+1}{2\sqrt{2}}\,,\,m \rightarrow \infty\,,
634: \label{largembehav}
635: \end{equation}
636: and indeed already for $m=4$ this formula yields very good agreement
637: with the results shown in Table \ref{tab:fundnonrot}. This very same
638: result can also be obtained using a P\"oschl-Teller fitting potential
639: \cite{ferrari}. Just to give an idea of the orders of magnitude
640: involved consider an $m=2$ mode: if we take the ``typical'' gravity
641: wave analogue experimental parameters of Section \ref{gwanal} we get
642: (for the fundamental QNM with $m=2$) a frequency $\omega_R=0.945
643: \times c^2/A=0.293$ Hz and a damping timescale $\tau=1/|\omega_I| =
644: 1/0.344 \times A/c^2=9.37$ s.
645:
646: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
647: \subsubsection{Slowly decaying modes of rotating black holes}
648: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
649:
650: For rotating black holes, the generalized potential has more than one
651: extremum. This immediately raises problems for the applicability of
652: the WKB technique. Previous work on the Kerr geometry
653: \cite{will,kokkotaswkb} showed that the WKB method can still be used,
654: yielding good results, as long as the rotation parameter is small.
655: The way to handle the several extrema of the potential is the
656: following: start with $B=0$ and compute the roots of $dQ/dr_*=0$ (to
657: find the extrema; as we said before, for $B=0$ the generalized
658: potential $Q$ has a single maximum outside the horizon as long as
659: $m\neq 0$). Compute $\omega _{QN}$ in the non-rotating case. Then
660: add some small rotation $B$ to the black hole, and compute the roots
661: of $dQ/dr_*=0$. Now the roots will depend on $\omega$, but only one
662: root yields the correct non-rotating limit as $B\rightarrow 0$. It is
663: this root that corresponds to the QNMs of the rotating black hole. Our
664: results, which can be trusted for $B\lesssim 0.2$, are shown in
665: Fig. \ref{fig:QNM.eps}. For $m>0$ ($m<0$) $\omega_R$ and $|\omega_I|$
666: increase (decrease) with rotation, at least in the range in which the
667: WKB method can be applied. The QN frequency change with rotation is
668: not dramatic, but it can probably be used to apply a ``fingerprint
669: analysis'' of the acoustic black hole parameters {\it \'a la
670: Echeverria} \cite{echeverria}: that is, once we measure at least two
671: QNM frequencies we may infer the acoustic black hole parameters $A$ and $B$.
672: Carrying out such experiments in the lab can shed some light on the
673: applicability of similar ideas to test the no-hair theorem in the
674: astrophysical context \cite{dreyerfinn}. A more complete analysis of
675: the QNMs is still needed to probe the high rotation regime: the
676: results shown in Fig. \ref{fig:QNM.eps} seem to indicate that
677: instability may set in for large $Bm$. A continued fraction analysis
678: could be used to test this hypothesis, but it is beyond the scopes of
679: this paper \cite{cardoso4}.
680: \vskip 1mm
681: \begin{figure}
682: \centerline{\mbox{
683: \psfig{figure=fig1a.ps,angle=270,width=9cm}
684: \psfig{figure=fig1b.ps,angle=270,width=9cm}
685: }}
686: \caption{ In the left panel we show the real part of the fundamental
687: QN frequency $\omega_R$ (and in the right panel we show the imaginary
688: part $\omega_I$) as a function of the rotation parameter $B$, for
689: selected values of $m$. For $m>0$ ($m<0$) $\omega_R$ and $|\omega_I|$
690: increase (decrease) with rotation. }
691: \label{fig:QNM.eps}
692: \end{figure}
693:
694: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
695: \subsubsection{Highly damped modes}
696: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
697:
698: QNMs with a large imaginary part, i.e. with a very large overtone
699: number $n$, have recently become a subject of intense scrutiny for
700: black holes in general relativity and similar theories. The interest
701: in these modes comes from Hod's proposal \cite{hod} that they could be
702: related to black hole area quantization, and from Dreyer's suggestion
703: that a similar argument could be used to fix the Barbero-Immirzi
704: parameter in Loop Quantum Gravity \cite{dreyer}. In this context, an
705: analytical calculation of highly damped QNMs was first carried out by
706: Motl for the Schwarzschild black hole \cite{motl1}. Subsequently Motl
707: and Neitzke \cite{motl2} used a complex-integration technique to
708: compute highly-damped QNMs of the Schwarzschild and
709: Reissner-Nordstr\"om black holes. Their analytical results are in
710: striking agreement with numerical data \cite{num1,num3} and
711: alternative analytical calculations \cite{nils}. The complex
712: integration method has also been generalized with success to other
713: black hole geometries \cite{ricardo,tamaki},
714: yielding again predictions in
715: excellent agreement with the numerical results \cite{num2}. In view of
716: this, we shall now build on the results and techniques of \cite{motl2}
717: to compute highly damped QNMs of a rotating acoustic black hole.
718:
719: Let us first consider a non-rotating black hole. In the limit $B\to
720: 0$ the wave equation (\ref{waveequation2}) reduces to
721: %
722: \be
723: \frac{d^2H}{dr_*^2}+
724: \left [ \omega ^2-
725: V(r)
726: \right] H=0\,,
727: \ee
728: where the potential
729: \be
730: V(r)\equiv \left ( 1-\frac{1}{r^2}\right )
731: \left (\frac{m^2-1/4}{r^2}+\frac{5}{4r^4} \right )\,.
732: \ee
733: %
734: Following (\cite{motl2}), we can determine the highly damped QNMs
735: looking at the behaviour of the potential near the singular point
736: $r=0$. In our case the leading term as $r\to 0$ is
737: %
738: \be
739: V \sim -\frac{5}{4r^6}\,,
740: \ee
741: and in the same limit we have $r_* \sim -r^3/3$. Thus
742: \be
743: V \sim -\frac{5}{36r_*^2}=\frac{j^2-1}{4r_*^2}\,,
744: \ee
745: %
746: for $j=2/3$ (this is precisely the same power-law behavior found in
747: the case of the Schwarzschild black hole, except for the different
748: value of $j$). The result in \cite{motl2} carries over directly:
749: %
750: \be
751: e^{4\pi \omega}=-(1+2\cos{\pi j})\,.
752: \ee
753: %
754: However now $(1+2\cos{\pi j})=0$. This means that there are no
755: asymptotic QN frequencies for this black hole. A similar situation
756: occurs for electromagnetic perturbations of a five-dimensional black
757: hole \cite{num2}. If we add rotation to the black hole, i.e., if $B$
758: is non-zero, a similar analysis \cite{neitzke} implies that the
759: asymptotic QN frequencies behave as
760: %
761: \be
762: e^{4\pi (\omega-mB)}=-(1+2\cos{\pi j})\,,
763: \ee
764: %
765: where again $j=2/3$. Thus also in this case there will be no
766: asymptotic QN frequencies.
767:
768: This result is quite puzzling. It could mean either that the real part
769: grows without bound, or that it just doesn't converge to any finite
770: value. In any case, an application of Hod's conjecture to these black
771: holes seems impossible: the very definition of the classical laws of
772: black hole thermodynamics is non-trivial in the analogue case. In
773: fact, as argued by Visser \cite{visserlaws}, the laws of black hole
774: thermodynamics arise solely from the Einstein equations for the
775: metric; from a different perspective, it has been shown that the
776: Einstein equations themselves can be derived from purely
777: thermodynamical arguments \cite{TedEinsteinEOS}. A well-known
778: ``weakness'' of analogue models is the fact that they can be used to
779: reproduce the kinematical aspects of the Einstein equations, but not
780: their dynamical aspects (see eg. \cite{BLSV} for the implications of
781: this distinction on the causal structure of acoustic
782: spacetimes). Hod's arguments \cite{hod} assume a thermodynamic
783: relation between black hole surface area and entropy: even if Hod's
784: conjecture is true, the absence of any such relation for analogues
785: could explain the missing link between the QNM spectrum and area
786: quantization.
787:
788: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
789: \subsection{Late-time tails}
790: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
791:
792: After the exponential QNM decay characteristic of the ringdown phase
793: black hole perturbations usually decay with a power-law tail
794: \cite{price1}, due to the backscattering of waves off the background
795: curvature. The existence of late-time tails in black hole spacetimes
796: is by now well established. A strong body of evidence comes from
797: analytical and numerical calculations using linear perturbation theory
798: and even non-linear evolutions, for massless or massive fields
799: \cite{price1,price2,ching1,ching2,cardosoDdimtails}. In a very
800: complete analysis, Ching, Leung, Suen and Young \cite{ching1,ching2}
801: considered the late-time tails appearing when one deals with evolution
802: equations of the form (\ref{waveequation2}), and the potential $V$ is
803: of the form
804: \begin{equation}
805: V(r_*) \sim \frac{\nu(\nu+1)}{r_*^2}+
806: \frac{c_1\log{r_*}+c_2}{r_*^{\alpha}}\,,\qquad r_*\rightarrow \infty.
807: \label{potching}
808: \end{equation}
809: By a careful study of the branch cut contribution to the associated
810: Green's function they concluded that in general the late-time behavior
811: is dictated by a power-law or by a power-law times a logarithm. The
812: exponents of the power-law depend on the leading term at very large
813: spatial distances. The case of interest for us here is when
814: $c_1=0$. Their conclusions, which we will therefore restrict to the
815: $c_1=0$ case, are (see Table 1 in \cite{ching1} or \cite{ching2}):
816: \newline
817: (i) if $\nu$ is an integer the term $\nu(\nu+1)/r_*^2$ does not
818: contribute to the late-time tail. Since this term represents just the
819: pure centrifugal barrier, characteristic of flat space, one can expect
820: that indeed it does not contribute, at least in four-dimensional
821: spacetimes. Therefore, for integer $\nu$, it is the $c_2/r_*^{\alpha}$
822: term that contributes to the late-time tail. In this case the authors
823: of \cite{ching1,ching2} find that the tail is given by a power-law,
824: \begin{equation}
825: \Psi \sim t^{-\mu}\,\,,\,\,\mu>2\nu+\alpha\,\,,\,\, \alpha\,
826: {\rm odd}\,\,{\rm integer}<2\nu+3.
827: \label{tailintnu1}
828: \end{equation}
829: where the exponent $\mu=2\nu+2\alpha-2$. For all other real $\alpha$,
830: the tail is
831: \begin{equation}
832: \Psi \sim t^{-(2\nu+\alpha)}\,\,,\,\,{\rm all}\,\,{\rm other}\,\,
833: {\rm real}\,\,\alpha.
834: \label{tailintnu2}
835: \end{equation}
836: \newline
837: (ii) if $\nu$ is not an integer, then the main contribution to the
838: late-time tail comes from the $\nu(\nu+1)/r_*^2$ term. In this case
839: the tail is
840: \begin{equation}
841: \Psi \sim t^{-(2\nu+2)}\,\,,\,\,{\rm non\,integer}\,\,\nu.
842: \label{tailnonintnu}
843: \end{equation}
844: As an example of non-integer $\nu$, Cardoso {\it et al.}
845: \cite{cardosoDdimtails} have shown that the late-time tails of wave
846: propagation appearing in odd dimensional spacetimes do not depend on
847: the presence of the black hole at all. For odd dimensions the
848: power-law is determined not by the presence of the black hole, but by
849: the very fact that the spacetime is odd dimensional. In this case the
850: field decays as
851: %
852: \be
853: \Psi \sim t^{-(2l+D-2)}\,,
854: \label{tailsD}
855: \ee
856: %
857: where $l$ is the angular index determining the angular dependence of
858: the field, and $D$ the number of spacetime dimensions. One can show
859: directly from the flat space Green's function that such a power-law is
860: indeed expected in flat, odd dimensional spacetimes.
861:
862: From the aforementioned arguments we should expect late-time tails in
863: our draining bathtub acoustic black hole to be directly related to the
864: dimensionality of the underlying flow, and not to the presence of a
865: black hole, since we are dealing with a $(2+1)$-dimensional
866: flow. Indeed, at leading order the potential behaves as
867: %
868: \be
869: V \sim \frac{m^2-1/4}{r_*^2}=\frac{\nu(\nu+1)}{r_*^2}\,,
870: \label{as1}
871: \ee
872: %
873: where $\nu=m-1/2$. We are thus in case (ii) above, because $\nu$ is
874: not an integer. So any perturbation in the vicinities of this black
875: hole will die out as a late-time tail of the form
876: %
877: \be
878: H \sim t^{-(2m+1)}\,.
879: \label{latet}
880: \ee
881: %
882: This power-law is just the one given by the general expression
883: (\ref{tailsD}), if one substitutes $D=3$.
884:
885: Strictly speaking, the asymptotic form (\ref{as1}) for the potential
886: is valid only for non-rotating acoustic black holes. For the general
887: rotating case the leading term in the potential is $\omega$-dependent:
888: %
889: \be
890: V \sim \frac{m^2-1/4}{r_*^2}=\frac{\nu(\nu+1)+2\omega Bm}{r_*^2}\,.
891: \label{as2}
892: \ee
893: %
894: Since late-times are associated with small frequencies, {\it and} the
895: term $\nu(\nu+1)$ gives a contribution to the late-time tail, then it
896: follows that the late-time tails of rotating acoustic black holes are
897: the same as those for non-rotating black holes, expression
898: (\ref{latet}).
899:
900: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
901: \subsection{Superradiance}\label{integration}
902: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
903:
904: Superradiance is a general phenomenon in physics. Inertial motion
905: superradiance has long been known \cite{ginzburg}, and refers to the
906: possibility that a (possibly electrically neutral) object endowed with
907: internal structure, moving uniformly through a medium, may emit
908: photons even when it starts off in its ground state. Some examples of
909: inertial motion superradiance include the Cherenkov effect, the Landau
910: criterion for disappearance of superfluidity, and Mach shocks for
911: solid objects travelling through a fluid (cf. \cite{bekschiffer} for a
912: discussion). Non-inertial rotational motion also produces
913: superradiance. This was discovered by Zel'dovich \cite{zeldovich}, who
914: pointed out that a cylinder made of absorbing material and rotating
915: around its axis with frequency $\Omega$ can amplify modes of scalar or
916: electromagnetic radiation of frequency $\omega$, provided the
917: condition
918: %
919: \be\label{suprad}
920: \omega<m\Omega
921: \ee
922: %
923: (where $m$ is the azimuthal quantum number with respect to the axis of
924: rotation) is satisfied. Zel'dovich realized that, accounting for
925: quantum effects, the rotating object should emit spontaneously in this
926: superradiant regime. He then suggested that a Kerr black hole whose
927: angular velocity at the horizon is $\Omega$ will show both
928: amplification and spontaneous emission when the condition
929: (\ref{suprad}) for superradiance is satisfied. This suggestion was put
930: on firmer ground by a substantial body of work \cite{superr}. In
931: particular, it became clear that (even at the purely {\it classical}
932: level) superradiance is required to satisfy Hawking's area theorem
933: \cite{beksuperr,pressteu}.
934:
935: Superradiance is essentially related to the presence of an ergosphere,
936: allowing the extraction of rotational energy from a black hole through
937: a wave equivalent of the Penrose process \cite{penrose}. Under certain
938: conditions, superradiance can be used to induce instabilities in Kerr
939: black holes \cite{schwinger}. Indeed, all spacetimes admitting an
940: ergosphere and {\it no horizon} are unstable due to rotational
941: superradiance. This was shown rigorously in \cite{friedman}, but the
942: growth rate of the instability is too slow to observe it in an
943: astrophysical context \cite{cominsschutz}. Kerr black holes are
944: stable, but if enclosed by a reflecting mirror they can become
945: unstable due to superradiance \cite{ptbhb,bhb}; we will discuss this
946: ``black hole bomb'' instability as applied to the analogue acoustic
947: black hole.
948:
949: The possibility to observe rotational superradiance in analogue black
950: holes was considered by Sch\"utzhold and Unruh \cite{schutzhold}, and
951: more extensively by Basak and Majumdar \cite{basak1,basak2}, who
952: computed analytically the reflection coefficients in the low frequency
953: limit $\omega A/c^2\ll 1$. In particular, the authors of
954: \cite{schutzhold} showed that the ergoregion instability in gravity
955: wave analogues is related to the existence of an ``energy function''
956: [their Eq. (68)] that is not positive definite inside the
957: ergosphere. In the context of analogues, inertial superradiance based
958: on superfluid $^3$He has been studied by Jacobson and Volovik
959: \cite{tedvolovik}.
960:
961: Here we present a quantitative calculation of the efficiency of
962: superradiant amplification for the draining bathtub metric. We work in
963: the frequency representation, and consider an incident plane wave of
964: unit amplitude at infinity, frequency $\omega$ and azimuthal index
965: $m$. Part of this wave will be reflected back by the medium, the
966: reflection coefficient being some (complex) number $R_{\omega m}$. In
967: terms of the wave equation (\ref{waveequation2}), this determines the
968: following boundary condition at infinity:
969: %
970: \be
971: H\sim R_{\omega m} e^{\ii \omega r_*}+e^{-\ii \omega r_*}
972: \,\,,r\rightarrow \infty\,.
973: \label{asymptbeh}
974: \ee
975: %
976: At the sonic horizon ($r\to 1$, $r_*\to -\infty$) the solution behaves
977: like
978: %
979: \be
980: H\sim T_{\omega m} e^{-\ii (\omega-mB) r_*}\,\,,r\rightarrow 1\,.
981: \label{asymptbeh1}
982: \ee
983: where $T_{\omega m}$ is the transmission coefficient.
984: %
985: An easy way to prove the existence of superresonance is to compute the
986: Wronskian of a solution of equation (\ref{waveequation2}) and of its
987: adjoint at the sonic horizon and at infinity. From the constancy of
988: the Wronskian as a function of the radial coordinate, using the
989: boundary conditions (\ref{asymptbeh}) and (\ref{asymptbeh1}) we find
990: the following ``energy conservation'' condition:
991: %
992: \be
993: 1-\left|R_{\omega m}\right|^2=
994: \left(1-\f{mB}{\omega}\right)\left|T_{\omega m}\right|^2\,.
995: \ee
996: %
997: Therefore, if $\omega <mB$ the reflection coefficient $\left|R_{\omega
998: m}\right|^2>1$, i.e. we have superradiance.
999:
1000: To compute numerically with improved accuracy the reflection
1001: coefficient $R_{\omega m}$, we have used a refined condition at the
1002: horizon:
1003: %
1004: \be\label{horizon}
1005: H\sim T_{\omega m} e^{-\ii (\omega-mB) r_*}\left[1+y_1(r-1)+\dots\right]\,,
1006: \ee
1007: %
1008: where the leading-order coefficient is
1009: %
1010: \be
1011: y_1=\frac{(1+2B^2)m^2-2\omega Bm+1}{ 2[\ii(mB-\omega )+1]}\,.
1012: \ee
1013: %
1014: We also keep higher-order terms to extract the reflection coefficient
1015: at infinity. Namely, for the outgoing wave we use an expansion of the
1016: form:
1017: %
1018: \be\label{outg}
1019: H\sim e^{\ii \omega r_*}\left[1+\frac{z_1}{r}+\frac{z_2}{r^2}+\dots\right]
1020: \,,
1021: \ee
1022: %
1023: where
1024: %
1025: \be
1026: z_1=\ii\frac{B+(4m^2-1)}{ 8\omega}\,,\qquad
1027: z_2=\ii z_1\left[\frac{Bm}{2}+\frac{(4m^2-9)}{16\omega}\right]\,,\nn
1028: \ee
1029: %
1030: and for the ingoing wave we use the complex conjugate of
1031: Eq. (\ref{outg}). As a check of our code we have reproduced results
1032: by Andersson {\it et al.} \cite{alp} for the superradiant
1033: amplification of a scalar field in the vicinities of a Kerr black
1034: hole. Our results are in perfect agreement with Fig. 1 of
1035: \cite{alp}. In particular, we find a maximum amplification coefficient
1036: given by $1-|R_{\omega m}|^2\simeq 0.2$ \% for $l=m=2$ scalar
1037: perturbations of a near-extremal extremal Kerr black hole (in early
1038: numerical work \cite{ptbhb,pressteu} the maximum amplification was
1039: found to be $1-|R_{\omega m}|^2\simeq 0.3$ \%).
1040:
1041: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1042: \begin{figure}
1043: \centerline{\mbox{
1044: \psfig{figure=fig2a.ps,angle=270,width=9cm}
1045: \psfig{figure=fig2b.ps,angle=270,width=9cm}
1046: }}
1047: \centerline{\mbox{
1048: \psfig{figure=fig2c.ps,angle=270,width=9cm}
1049: \psfig{figure=fig2d.ps,angle=270,width=9cm}
1050: }}
1051: \centerline{\mbox{
1052: \psfig{figure=fig2e.ps,angle=270,width=9cm}
1053: \psfig{figure=fig2f.ps,angle=270,width=9cm}
1054: }}
1055: \caption{ Reflection coefficient $\left|R_{\omega m}\right|^2$ as a
1056: function of $\omega$ for $m=1$ (left panels) and $m=2$ (right
1057: panels). Each curve corresponds to a different value of $B$, as
1058: indicated. The top panels show that the reflection coefficient decays
1059: exponentially at the critical frequency for superradiance,
1060: $\omega_{SR}=mB$. The middle panels show a close-up view in the
1061: superradiant regime for $B<1$: at $B=1$ the maximum amplification is
1062: 21.2 \% ($m=1$) and 4.7 \% ($m=2$). The bottom panels show that
1063: superradiant amplification can become much more efficient for values
1064: of the rotation parameter $B>1$. }
1065: \label{fig:superradiantfactor}
1066: \end{figure}
1067: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1068:
1069: Results of the numerical integrations for the draining bathtub metric
1070: are shown in Fig. \ref{fig:superradiantfactor}. Panels on the left
1071: show the reflection coefficient $|R_{\omega 1}|^2$ for $m=1$, and
1072: panels on the right show $|R_{\omega 2}|^2$ for $m=2$, for selected
1073: values of the black hole rotation $B$. Panels on top show that, as
1074: expected, in the superradiant regime $0<\omega<mB$ the reflection
1075: coefficient $|R_{\omega m}|^2\geq 1$. Furthermore, as one increases
1076: $B$ the reflection coefficient increases, and for fixed $B$, the
1077: reflection coefficient $\left |R_{\omega m}\right|^2$ attains a
1078: maximum at $\omega \sim mB$, after which it decays exponentially as a
1079: function of $\omega$ outside the superradiant interval. This is very
1080: similar to what happens when one deals with massless fields in the
1081: vicinities of rotating Kerr black holes \cite{pressteu}. In particular,
1082: from the close-up view in the middle panels we see that, for $B=1$,
1083: the maximum amplification is 21.2 \% ($m=1$) and 4.7 \% ($m=2$).
1084:
1085: As a final remark, and as we have anticipated, an important difference
1086: between the acoustic black hole metric and the Kerr metric is that in
1087: the present case there is no mathematical upper limit on the black
1088: hole's rotational velocity $B$. In the bottom panels we show that,
1089: considering values of $B>1$, we can indeed have larger amplification
1090: factors for acoustic black holes.
1091:
1092: Summarizing: if we are clever enough to build in the lab an acoustic
1093: black hole that spins very rapidly, rotational superradiance can be
1094: particularly efficient in analogues. This is an important result,
1095: considering that the detection of rotational superradiance in the lab
1096: is by no means an easy task, as originally predicted by Zel'dovich
1097: \cite{zeldovich} and confirmed by recent reconsiderations of the
1098: problem \cite{bekschiffer}. Of course, in any real-world experiment
1099: the maximum rotational parameter will be limited. At the mathematical
1100: level, the equations describing sound propagation (which are written
1101: assuming the hydrodynamic approximation) will eventually break
1102: down. Physically, if the angular component of the velocity $v^\theta$
1103: becomes very large the dispersion relation for the fluid will change,
1104: invalidating the assumptions under which we have derived our acoustic
1105: metric \cite{schutzhold}. To be more concrete, let us consider again
1106: the gravity wave analogue described in Section \ref{gwanal}. Then we
1107: can use a very simple argument to limit the acoustic black hole
1108: spin. The derivation of the rotating acoustic black hole metric is
1109: based on the assumption that the bottom of the tank should not be too
1110: steep, that is, $f'(r)\ll 1$. This condition translates into a
1111: condition for $B$:
1112: %
1113: \be\label{Blimit}
1114: B\ll\left(\frac{gr^3}{2}-A^2\right)^{1/2}\,,
1115: \ee
1116: %
1117: which must be satisfied for all values of $r$, and in particular at
1118: the acoustic black hole horizon $r=r_H$ (notice that here, and here
1119: only, we have switched back to physical units). Setting $r=r_H$ in the
1120: previous inequality and using the physical parameters quoted in
1121: Section \ref{gwanal} we get the very stringent condition that $B\ll
1122: 2.2$ m$^2$~s$^{-1}$, or (in the dimensionless units we use throughout
1123: this paper) $\hat B\ll 7.0\,$. In other words, we can only get values
1124: of the rotation parameter larger than $\hat B\sim 1$ when the slope of
1125: the tank is so large that the assumptions underlying the derivation of
1126: the acoustic metric are not valid any more. Of course, this example
1127: does not mean that such a constraint applies to {\it every} possible
1128: experimental realization of the draining bathtub metric. However, it
1129: serves as an illustration of the kind of experimental difficulties we
1130: may expect to encounter in practice.
1131:
1132: The superradiant phenomena we have described are purely {\it
1133: classical} in nature. However, an interesting suggestion to observe
1134: {\it quantum} effects in acoustic superradiance was put forward in
1135: \cite{basak2}. To write down our acoustic metric we required the flow
1136: to be irrotational and nonviscous. As a natural choice, we could use a
1137: fluid which is well known to possess precisely these properties:
1138: superfluid HeII. In this case the presence of vortices with quantized
1139: angular momenta may lead to a quantized energy flux. The heuristic
1140: argument presented in \cite{basak2} goes as follows. Let us imagine
1141: that our black hole is a vortex with a sink at the centre. In the
1142: quantum theory of HeII the wavefunction is of the form
1143: $\Psi=\exp\left[\ii \sum_j \phi(\vec r_j) \Phi_{\rm ground}\right]$,
1144: where $\vec r_j$ is the position of the $j$-th particle of HeII. The
1145: velocity at any point is given by the gradient of the phase at that
1146: point, $\vec v=\nabla \phi$, so that (roughly speaking) the velocity
1147: potential (\ref{vpot}) can be identified with the phase of the
1148: wavefunction. This phase will be singular at the sink
1149: $r=0$. Continuity of the phase around a circle surrounding the sink
1150: requires that the change of the wavefunction satisfies $\Delta
1151: \phi=2\pi B$. For the wavefunction to be single valued, $B$ (that is,
1152: the black hole's angular velocity at the horizon) must be the integer
1153: multiple of some minimum value $\Delta B$, i.e., $B=n\,\Delta B$. Then
1154: the angular momentum of the acoustic black hole would be forced to
1155: change in integer multiples of $\Delta B$. Correspondingly, the
1156: spectrum of the reflection coefficients may be given by equally-spaced
1157: peaks with different strengths. This discrete amplification could
1158: enhance chances of observing superradiance in acoustic black holes,
1159: and rule out (or provide empirical support to) some of the many
1160: competing heuristic approaches to black hole quantization.
1161:
1162: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1163: \subsection{Superradiant instabilities: the acoustic black hole bomb}
1164: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1165:
1166: The very existence of an ergoregion in the acoustic black hole metric
1167: allows immediately for the possibility to make the system
1168: unstable. Suppose we enclose the system by a reflecting mirror at
1169: constant radius $r_0$. Now, throw in a wave having frequency $\omega=
1170: \omega_R+{\rm i}\, \omega_I$ such that $\omega_R<m\Omega$: the wave
1171: will be amplified at the expense of the black hole's rotational energy
1172: and travel back to the mirror. There it will be reflected and move
1173: again towards the black hole, this time with increased
1174: amplitude. Through repeated reflections, the waves' amplitude will
1175: grow exponentially with time. This ``black hole bomb'' was first
1176: proposed by Press and Teukolsky \cite{ptbhb}, and recently it has been
1177: studied in detail by some of us \cite{bhb} in the context of the Kerr
1178: metric. The detailed analysis performed in \cite{bhb} showed that the
1179: black hole bomb can be characterized by a set of complex resonant
1180: frequencies, the Boxed QuasiNormal Modes (BQNMs). The real part of a
1181: BQNM is, not surprisingly, proportional to $1/r_0$, where $r_0$ is the
1182: mirror radius: this is essentially the condition for the existence of
1183: standing waves in the region enclosed by the mirror. The imaginary
1184: part of the BQNMs is proportional to $(\omega_R-m\Omega)$: the system
1185: can only become unstable if it is in the superradiant regime. Combined
1186: with the standing wave condition this implies that the mirror should
1187: be placed at some radius $r_0\gtrsim 1/m\Omega$ in order for the
1188: system to be unstable. As rotational energy is extracted from the
1189: system the black hole will spin down. For any given $r_0$ the
1190: instability will eventually shut off, as the condition $r_0\gtrsim
1191: 1/m\Omega$ will no longer be satisfied.
1192:
1193: An additional reason to study the system ``acoustic black
1194: hole+mirror'' is related to the problem of boundary conditions at
1195: infinity for QNMs (cf. Section \ref{dbqnms}). Whatever analogue model
1196: we consider, no physical apparatus will ever extend out to
1197: infinity. We are left with two possibilities: we can either use some
1198: absorbing device to simulate spatial infinity, or we can impose
1199: alternative boundary conditions on the system. A natural choice is to
1200: have a reflecting mirror surrounding the apparatus, so that we are
1201: effectively building an acoustic black hole bomb.
1202:
1203: In the following we will assume that we are in the presence of a
1204: perfectly reflecting mirror. Correspondingly, we shall impose the
1205: boundary condition $H=0$ at $r=r_0$. At the horizon we demand, as
1206: usual, the presence of purely ingoing waves (waves headed towards the
1207: horizon), i.e. $H \sim e^{-\ii(\omega-Bm)r_*}$. We are again in
1208: presence of an eigenvalue problem. However, since the boundary
1209: conditions have changed we shall not refer to the characteristic
1210: frequencies as QN frequencies, but rather as Boxed QuasiNormal
1211: frequencies (BQN frequencies). The reader is referred to \cite{bhb},
1212: where this terminology was first introduced. Using a direct
1213: integration of the wave equation (see Section
1214: \ref{integration} for details) it is quite easy to compute the BQN
1215: frequencies. We simply integrate Eq. (\ref{waveequation2}) outwards
1216: from the horizon, where we impose the condition (\ref{horizon}), to
1217: the mirror location $r_0$. The (complex) BQNM frequencies are those
1218: frequencies for which
1219: %
1220: \be
1221: H(\omega_{BQNM},r_0)=0\,.
1222: \ee
1223: %
1224: Some results for the fundamental BQNM and the first two overtones are
1225: shown in Fig. \ref{fig:BQNM1}. There we fixed the black hole
1226: rotation parameter $B=1$, and $m=1$, as a typical case.
1227:
1228: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1229: \begin{figure}
1230: \centerline{\mbox{
1231: \psfig{figure=fig3a.ps,angle=270,width=9cm}
1232: \psfig{figure=fig3b.ps,angle=270,width=9cm}}}
1233: \caption{ Real part (left) and imaginary part (right) of the
1234: fundamental BQN frequency and the first two overtones as a function of
1235: mirror location $r_0$. Both panels refer to $m=1$ and $B=1$ (but the
1236: $B$-dependence of the real part of the BQN frequency is very weak).
1237: With excellent accuracy $\omega_I$ crosses zero (and the instability
1238: abruptly shuts off) at the critical radius predicted by
1239: Eq. (\ref{bomboff}), as can be verified taking a glance at the first
1240: row of Table \ref{tab:bessel}. }
1241: \label{fig:BQNM1}
1242: \end{figure}
1243: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1244:
1245: From the left panel we see that the real part of the BQNM frequency
1246: for the $n$-th overtone (where we use the convention that $n=0$
1247: corresponds to the fundamental mode) scales as $1/r_0$, as
1248: anticipated. The proportionality constant can be obtained by
1249: analytical arguments similar to those in \cite{bhb}. The result is
1250: %
1251: \beq\label{BQNMr}
1252: \omega_R=\frac{j_{m,n}}{r_0},
1253: \eeq
1254: %
1255: where the $j_{m,n}$'s are zeros of the Bessel function of {\it
1256: integer} order $m$ (recall that for a Kerr black hole bomb one gets
1257: $\omega_R=j_{l+1/2,n}/r_0$ instead, cf. \cite{bhb}). The numerical
1258: values of the $j_{m,n}$'s can be found, e.g., in \cite{abramowitz}. For
1259: reference, we list the first few values in Table \ref{tab:bessel}.
1260: Our numerical calculations are in excellent agreement with the
1261: predictions of formula (\ref{BQNMr}). This is particularly true for
1262: small values of $B$, but the $B$-dependence of the real part of the
1263: BQN frequencies is very weak anyway, as it is for the
1264: Kerr metric \cite{bhb}.
1265:
1266: From the right panel we see the behavior predicted by the qualitative
1267: arguments at the beginning of this Section: for any BQNM, the
1268: instability is more and more efficient as the mirror radius $r_0$
1269: becomes smaller, until eventually the mirror radius becomes small
1270: enough that the instability shuts off.
1271: %For overtones, a simple empirical rule to get the order of magnitude
1272: %of the oscillation frequency is: $\omega_R\simeq (j_{n,0}+n\pi)/r_0$.
1273:
1274: \begin{table}
1275: \centering
1276: \caption{\label{tab:bessel} Zeros of the Bessel functions $j_{m,n}$
1277: for the first few values of $m$. The real part of the BQNM frequency
1278: is well approximated by Eq. (\ref{BQNMr}).}
1279: \vskip 12pt
1280: \begin{tabular}{@{}c|c|c|c@{}}
1281: \hline
1282: \hline
1283: $m$ &$j_{m,0}$ &$j_{m,1}$ &$j_{m,2}$\\
1284: \hline
1285: \hline
1286: 1 &3.83171 &7.01559 &10.17347\\
1287: 2 &5.13562 &8.41724 &11.61984\\
1288: 3 &6.38016 &9.76102 &13.01520\\
1289: 4 &7.58834 &11.06471 &14.37254\\
1290: 5 &8.77148 &12.33860 &15.70017\\
1291: \hline
1292: \hline
1293: \end{tabular}
1294: \end{table}
1295:
1296: A few remarks are in order: (i) The real part of the BQN frequency
1297: (slowly) increases with overtone number $n$. (ii) Higher overtones become
1298: stable at larger distances, but they also attain a smaller maximum
1299: growing rate. (iii) With excellent accuracy, the instability switches
1300: off at the critical radius predicted by the analytical formula
1301: (\ref{BQNMr}) supplemented by the superradiance condition
1302: (\ref{suprad}), that is:
1303: %
1304: \be
1305: r_{0,c}\simeq \frac{j_{m,n}}{mB}\,.
1306: \label{bomboff}
1307: \ee
1308: %
1309: Since $j_{m,n}$ grows linearly with $m$ (in particular, this is
1310: asymptotically true for high $m$'s \cite{abramowitz}) the critical
1311: radius is almost $m$-independent. However, the instability is not so
1312: efficient for higher $m$ as it is for small $m$, at least when
1313: $B\lesssim 1$. This is apparent in Fig. \ref{fig:BQNM2}, where we
1314: compare growth timescales with $m=1$ (left) and $m=2$ (right) for
1315: acoustic black holes rotating at different rates. For example, the
1316: maximum growth rate for $B=0.3$ is $\sim 10^{-4}$ for $m=1$, and $\sim
1317: 10^{-5}$ for $m=2$. The growth rate for $m=2$ becomes roughly
1318: comparable to the growth rate for $m=1$ when $B>1$: look for example
1319: at the curves corresponding to $B=5.0$. However, other physical
1320: considerations may forbid the construction of acoustic black holes
1321: rotating at such rates (see the arguments at the end of Section
1322: \ref{integration}).
1323:
1324: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1325: \begin{figure}
1326: \centerline{\mbox{
1327: \psfig{figure=fig4a.ps,angle=270,width=9cm}
1328: \psfig{figure=fig4b.ps,angle=270,width=9cm}}}
1329: \caption{
1330: Growth timescales for BQNMs, as a function of mirror location $r_0$,
1331: for different values of $B$. The left panel refers to $m=1$, the right
1332: panel to $m=2$. Notice the different scales for the $\omega_I$ axis.
1333: }
1334: \label{fig:BQNM2}
1335: \end{figure}
1336: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1337:
1338: To get an idea of the orders of magnitude involved, let us consider
1339: again the ``typical'' parameters for the gravity wave analogue we
1340: introduced in Section \ref{gwanal}. Let us pick $m=1$ and an acoustic
1341: black hole rotation parameter $\hat B=1$, which -- according to the
1342: arguments in Section \ref{integration} -- is close to the maximum
1343: rotation rate we may hope to achieve (recall hats distinguish
1344: dimensionless quantities from quantities in physical units). To take
1345: advantage of the process the mirror should be located close to the maximum
1346: of $\omega_I$, but not quite at the maximum. For example, if we place
1347: the mirror at $\hat r_0\simeq 10$ (cf. the right panel of
1348: Fig. \ref{fig:BQNM1}) we get $\hat \omega_I\simeq 4\cdot 10^{-3}$ (in
1349: this case the maximum growth rate would be $\hat \omega_I=8.5 \cdot
1350: 10^{-3}$). This corresponds to a growth time $\tau\simeq 800\,{\rm
1351: s}\simeq 13\,{\rm minutes}$ (at the maximum we would get $\tau\simeq
1352: 6$~minutes). Roughly speaking, this means that the amplitude will
1353: double every 13 minutes or so, and will be amplified by orders of
1354: magnitude on timescales of the order of a few hours. To reduce the
1355: typical timescale to seconds, one just has to ajust the horizon radius
1356: and wave velocity. For example, working with $r_H \sim 0.1\,$m and
1357: $c\sim 10$ m~s$^{-1}$, we would get a typical timescale of about 2
1358: seconds. This looks like a perfectly reasonable timescale to observe
1359: an acoustic black hole bomb in the lab.
1360:
1361: When the instability sets in the acoustic black hole loses energy and
1362: angular momentum ($\Delta E\sim B \Delta J$). The critical radius
1363: $r_{0,c}$ increases with decreasing $B$. To estimate the efficiency we
1364: can use a simple argument. We can take advantage of the mechanism by
1365: picking some large $B$ and $r_0\simeq 2 r_{0, max}$, where $r_{0,
1366: max}$ is the radius of maximum growth timescale at the given $B$. As
1367: the hole emits angular momentum $B$ will decrease. For any given
1368: mirror location $r_0$, the bomb eventually switches off when condition
1369: (\ref{bomboff}) is satisfied. From the difference in initial and final
1370: angular momentum $\Delta B$ we can infer the extracted energy $\Delta
1371: E$. For more details we refer to \cite{bhb}.
1372:
1373: As a final remark we want to stress that, when we talk about
1374: surrounding the acoustic black hole by a reflecting mirror, what we
1375: have in mind is a generic type of mirror. For instance, in the
1376: shallow basin gravity wave analogue, the mirror could be a circular
1377: rubber band of radius $r_0$ reflecting gravity waves. A possible
1378: alternative to implement a mirror could involve, for example,
1379: variations in the dispersion relation for sound waves. In particular,
1380: Sch\"utzhold and Unruh \cite{schutzhold} suggested that a simple
1381: change of the height $h_\infty$ of the basin would imply a change of
1382: the gravity wave speed, hence of the effective refractive index. The
1383: changed refractive index may be used to trap the waves, which can then
1384: be amplified via superradiant scattering until non-linear effects
1385: dominate.
1386:
1387: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1388: \section{The canonical non-rotating acoustic black hole}\label{canonical}
1389: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1390:
1391: In the previous Section we considered the metric describing a rotating
1392: $(2+1)$-dimensional acoustic black hole. Rotation implies the presence
1393: of an ergosphere, allowing us to study in the lab some of the most
1394: interesting phenomena concerning black hole physics (eg.
1395: superradiance and the related rotational instabilities). However, the
1396: non-rotating limit of the acoustic metric we considered is not the
1397: ``natural'' metric describing a non-rotating acoustic black hole. In
1398: this Section we are going to consider another class of non-rotating,
1399: $(3+1)$-dimensional acoustic metrics. These metrics are associated to
1400: the most general spherically symmetric flow of an incompressible
1401: fluid: in this sense they represent ``canonical'' metrics for a
1402: non-rotating acoustic black hole.
1403:
1404: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1405: \subsection{Formalism and basic equations}
1406: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1407:
1408: It is simple to show that the acoustic metric corresponding to the
1409: most general spherically symmetric flow of an incompressible fluid is
1410: given by \cite{visser}
1411: %
1412: \begin{equation}
1413: ds^2=
1414: -c^2\left (1-\frac{r_0^4}{r^4} \right)dt^2+
1415: \left (1-\frac{r_0^4}{r^4} \right)^{-1} dr^2+
1416: r^2\left (d\theta ^2+ \sin{\theta}^2 d\phi^2 \right)\,.
1417: \label{metriccanonical}
1418: \end{equation}
1419: %
1420: This metric does not correspond to any of the geometries typically
1421: considered in general relativity, but it describes (in the sense
1422: specified above) a ``canonical'' acoustic black hole. The propagation
1423: of small disturbances (sound waves) is again described by the massless
1424: Klein-Gordon equation $\nabla_{\mu}\nabla^{\mu}\Psi=0$ in this
1425: background. We can separate variables by the substitution
1426: %
1427: \be
1428: \Psi(t,r,\phi)=\frac{\Phi(\omega,r)}{r} e^{-\ii\omega t}Y_{lm}(\theta)\,,
1429: \ee
1430: %
1431: where $Y_{lm}(\theta)$ are the usual spherical harmonics. This yields
1432: the wave equation
1433: %
1434: \be
1435: \Phi_{,r_* r_*}+\left( \frac{\omega^2}{c^2}-V \right)\Phi=0\,,
1436: \label{waveequation4}
1437: \ee
1438: %
1439: where
1440: %
1441: \be
1442: V =\left ( 1-\frac{r_0^4}{r^4}\right )
1443: \left [ \frac{l(l+1)}{r^2}+\frac{4r_0^4}{r^6}\right ] \,,
1444: \label{pot5}
1445: \ee
1446: %
1447: and the tortoise coordinate $r_*$ is defined, as usual, by
1448: $dr/dr_*=(1-r_0^4/r^4)$. Since this is a spherical symmetric problem,
1449: the azimuthal number $m$ does not play any role, only the angular
1450: momentum $l$ is important here. In the next Section we summarize our
1451: results for QNMs and wave tails of the canonical acoustic black
1452: hole. When presenting our numerical results we will choose units such
1453: that $c=r_0=1$. This is of course equivalent to a simple rescaling of
1454: the radial variable ($\hat r=r/r_0$) and of the frequency ($\hat
1455: \omega=\omega/c$), but we will omit hats in the following.
1456:
1457: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1458: \subsection{Quasinormal modes}
1459: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1460:
1461: A 6th order WKB analysis \cite{konoplyawkb} reveals an unstable QNM
1462: for $l=0$. However, it is possible to prove stability of the canonical
1463: black hole since the potential is positive-definite. We therefore
1464: computed the QN frequencies using first the lowest approximation
1465: \cite{willwkb}, then 3rd order corrections \cite{will} and finally 6th
1466: order corrections. The results are presented in Table
1467: \ref{tab:fundnonrotcanonical}. The QN frequencies for $l=0$ and $l=1$
1468: seem to be the problem here. For any other $l$ the value quickly
1469: converges as we increase the correction order of the WKB
1470: method. Notice for example that the $l=0$ mode suffers a variation of
1471: almost an order of magnitude as we go from the lowest approximation to
1472: the 3rd order correction scheme, and gets unstable for the scheme
1473: using 6th order corrections. The reason for this failure is most
1474: likely related to the breakdown of the basic WKB assumptions, namely
1475: that the ratio of the derivatives of the potential to the potential
1476: itself is small. Indeed a close inspection shows that as $l$ increases
1477: these ratios tend to decrease. This means that the WKB method is more
1478: reliable for higher $l$: this was first observed in the early works on
1479: the subject \cite{willwkb,will}.
1480: For $l=0$ and $l=1$ the lowest WKB approximation gives the most
1481: reliable results. This is confirmed by recent numerical work \cite{num3}.
1482: \begin{table}
1483: \centering
1484: \caption{\label{tab:fundnonrotcanonical} The fundamental ($n=0$) QN
1485: frequencies for the non-rotating canonical acoustic black hole, using
1486: three WKB computational schemes. $\omega _{QN}^{(1)}$ is the result
1487: for the QN frequency using only the lowest approximation
1488: \cite{willwkb}, $\omega _{QN}^{(3)}$ is the value obtained using 3rd
1489: order improvements \cite{will}, and finally $\omega _{QN}^{(6)}$ was
1490: computed using 6th order corrections \cite{konoplyawkb}. }
1491: \vskip 12pt
1492: \begin{tabular}{@{}c|c|c|c@{}}
1493: \hline
1494: \hline
1495: $l$ &$\omega _{QN}^{(1)}$ &$\omega _{QN}^{(3)}$ &$\omega _{QN}^{(6)}$\\
1496: \hline
1497: \hline
1498: 0 & 1.13-0.73i & 0.19-1.11i &0.06+0.87i \\
1499: 1 & 1.37-0.68i & 0.53-0.71i &1.09-0.39i \\
1500: 2 & 1.82-0.64i & 1.41-0.61i &1.41-0.70i \\
1501: 3 & 2.36-0.62i & 2.10-0.62i &2.12-0.62i \\
1502: 4 & 2.94-0.62i & 2.75-0.62i &2.75-0.62i \\
1503: \hline
1504: \hline
1505: \end{tabular}
1506: \end{table}
1507:
1508: In the limit of large $l$ one finds
1509: %
1510: \be
1511: \omega \sim \frac{l\sqrt{2}}{3^{3/4}}-\ii\frac{\sqrt{2}(1+2n)}{3^{3/4}}\,,
1512: \ee
1513: %
1514: a result which agrees very well with our WKB data already for $l=3$.
1515:
1516: The calculation of highly damped QNMs proceeds along the same lines
1517: sketched for the $(2+1)$-dimensional acoustic black hole. In this case
1518: the index $j=3/5$, which implies that
1519: %
1520: \be
1521: 4\pi \omega =\log{\frac{3-\sqrt{5}}{2}} -\ii(2n+1)\pi\,.
1522: \ee
1523: %
1524: Once again asymptotic QN frequencies are not given by the logarithm of
1525: an integer, as required in Hod's construction \cite{hod}. As we
1526: remarked earlier this does not necessarily imply that the conjecture
1527: is wrong, since a thermodynamical interpretation of the black hole
1528: area is possible only when the dynamics of the system are described by
1529: the Einstein equations.
1530:
1531: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1532: \subsection{Late-time tails}
1533: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1534:
1535: The analysis of late-time tails proceeds as in the case of the
1536: $(2+1)$-dimensional acoustic black hole. We can easily show that
1537: asymptotically the potential behaves as
1538: %
1539: \be
1540: V \sim \frac{l(l+1)}{r_*^2}+\frac{12-5l(l+1)}{3r_*^6}\,,r\rightarrow
1541: \infty\,.
1542: \ee
1543: %
1544: Following the previous analysis we now find $\alpha=6$, and since $l$
1545: is an integer the power-law falloff is of the form
1546: %
1547: \be
1548: \Phi \sim t^{-(2l+6)}\,.
1549: \ee
1550: %
1551: Thus any perturbation eventually dies off as $t^{-(2l+6)}$, much more
1552: quickly (for $m=l$) than in the $(2+1)$-dimensional acoustic black
1553: hole background.
1554:
1555: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1556: \section{Conclusions}\label{conclusions}
1557: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1558:
1559: In this paper we have considered two acoustic black hole metrics: the
1560: $(2+1)$-dimensional ``draining bathtub'' metric, and the
1561: $(3+1)$-dimensional canonical non-rotating acoustic black hole. We
1562: have studied QNMs and late-time tails in both metrics, and
1563: superradiance in the ``draining bathtub'' metric.
1564:
1565: More specifically, we numerically computed slowly damped QNMs of the
1566: draining bathtub metric using a third-order WKB approach for all
1567: values of $m\neq 0$. We analytically evaluated QN frequencies in the
1568: large-$m$ limit [Eq. (\ref{largembehav})] and set upper limits on
1569: frequencies of unstable modes (Appendix \ref{unstable}). We showed
1570: that highly damped modes do not tend to any simple limit. At late
1571: times, power-law tails decay as $t^{-(2m+1)}$. This behavior is
1572: typical of any odd-dimensional spacetime: independently of the
1573: presence of a black hole, if $D$ is odd the power-law falloff is
1574: proportional to $t^{-(2l+D-2)}$ \cite{cardosoDdimtails} [the
1575: $(2+1)$-dimensional case, $D=3$, is special: the azimuthal and angular
1576: numbers are the same, $l=m$, since there is only one angular
1577: coordinate]. The draining bathtub metric, possessing an ergoregion,
1578: can superradiantly amplify waves. We computed reflection coefficients
1579: for this superradiant scattering by numerical integration. When the
1580: (dimensionless) acoustic black hole rotation $B=1$, the maximum
1581: amplification is 21.2 \% for $m=1$ and 4.7 \% for $m=2$. Enclosing
1582: the acoustic black hole by a reflecting mirror we can destabilize the
1583: system, making an initial perturbation grow exponentially with time:
1584: we have an acoustic black hole bomb \cite{pressteu}, or ``dumb hole
1585: bomb''. We computed analytically and numerically the frequencies and
1586: growing timescales for this instability. An interesting feature of
1587: acoustic geometries is that the acoustic black hole {\it spin} can be
1588: varied independently of the black hole {\it mass}. Therefore, at
1589: variance with the Kerr metric, the spin can be made (at least in
1590: principle) arbitrarily large, and rotational superradiance in acoustic
1591: black holes can be very efficient. Gravity waves in shallow water
1592: \cite{schutzhold} provide a concrete example of an experimental setup
1593: for studying classical physics in a draining bathtub metric [and of
1594: experimental difficulties in increasing arbitrarily the rotation rate:
1595: see the discussion leading to Eq. (\ref{Blimit})]. In this case, and
1596: for a typical choice of parameters, growth times for the black hole
1597: bomb instability would be of the order of minutes: this appears to be
1598: well within the range of experimental possibilities. Finally one can
1599: speculate, following \cite{basak2}, that superfluid HeII could be used
1600: to observe some quantum effects, such as a discrete superradiant
1601: amplification.
1602:
1603: For the canonical acoustic black hole we computed slowly-damped QNMs
1604: using first, third and sixth order WKB. For $l<2$ the WKB method does
1605: not seem to converge, and even yields an unstable frequency at 6th
1606: order when $l=0$. This is only due to bad convergence properties of
1607: the WKB technique, since the canonical acoustic black hole is
1608: stable. In the high-damping limit QNMs are given by $4\pi\omega =
1609: \log{[(3-\sqrt{5})/2]}-\ii(2n+1)\pi$, so they are not proportional to
1610: the logarithm of an integer, as required by recent conjectures.
1611: This is not surprising since there are no Einstein equations
1612: for acoustic metrics, i.e., the acoustic metric evolution
1613: is not governed by Einstein's equations.
1614: Therefore, acoustic black holes can teach us a lot about quantum gravity
1615: \cite{visser,corleyjacobson}, but cannot shed any light on its possible
1616: connection with highly damped quasinormal modes.
1617:
1618: The late time falloff in the canonical acoustic black hole metric
1619: is proportional to $t^{-(2l+6)}$,
1620: to be compared with the $t^{-(2l+3)}$ decay of 4-dimensional
1621: Schwarzschild black holes and with the $t^{-(2l+3D-8)}$ decay of
1622: even-dimensional spherically symmetric black holes with $D>4$
1623: \cite{cardosoDdimtails}.
1624:
1625: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1626: \section*{Acknowledgements}
1627: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1628:
1629: We are grateful to Carlos Barcel\'o and Ted Jacobson for useful
1630: discussions. This work was partially funded by Funda\c c\~ao para a
1631: Ci\^encia e Tecnologia (FCT) -- Portugal through project
1632: CERN/FNU/43797/2001. V.C. acknowledges financial support from FCT
1633: through grant SFRH/BPD/2003.
1634: This work was supported in part by the National Science Foundation under
1635: grant PHY 03-53180.
1636:
1637: \appendix
1638:
1639: \section{Upper bounds for quasinormal frequencies of unstable modes}
1640: \label{unstable}
1641:
1642: To derive some bounds on the magnitude of QN frequencies of possible
1643: unstable QNMs we shall follow Detweiler and Ipser \cite{det}. Let us
1644: begin with the Klein-Gordon equation for the evolution of the field,
1645: $\nabla_{\mu}\nabla^{\mu}\Psi=0$. Using the metric (\ref{metric2})
1646: and performing a mode decomposition $\Psi(t,r,\phi)=R(r)e^{\ii
1647: (m\phi-\omega t)}$, this can be written as
1648: %
1649: \be
1650: \frac{r}{fc^2}\omega ^2 R+\frac{d}{dr}\left[ rfR'\right]-
1651: \frac{2Bm\omega}{c^2fr}R-
1652: \left (\frac{1}{r}-\frac{B^2}{c^2r^3f}\right )m^2R=0\,,
1653: \label{eq1}
1654: \ee
1655: %
1656: where we have defined $f\equiv \Delta^{-1}=1-A^2/c^2r^2$.
1657:
1658: Multiply by $R^*$ and integrate from the horizon to spatial
1659: infinity. The result is
1660: %
1661: \be
1662: \int_{r_H} ^{\infty}dr \left[ \frac{r}{fc^2}\omega ^2 |R| ^2-rf|R'| ^2-
1663: \frac{2Bm\omega}{c^2fr}|R| ^2-
1664: \left (\frac{1}{r}-\frac{B^2}{c^2r^3f}\right )m^2|R|^2 \right ]=0\,,
1665: \label{eq2}
1666: \ee
1667: %
1668: where we have used an integration by parts and discarded the surface
1669: integrals. In fact, for unstable modes (and for these only) the
1670: boundary conditions guarantee that an unstable mode vanishes
1671: exponentially as $r \rightarrow r_H$ or $r\rightarrow \infty$. The
1672: imaginary part of this equation yields
1673: %
1674: \be
1675: \int_{r_H} ^{\infty}dr \left( r^2 \omega_R -Bm
1676: \right )\frac{|R| ^2}{frc^2}=0\,,
1677: \label{eq3}
1678: \ee
1679: %
1680: where we used $\omega=\omega_R+{\rm i}\,\omega_I$.
1681: Therefore $\omega_R$ and $Bm$ must have the same sign for unstable
1682: modes. Furthermore, since $|R| ^2/frc^2$ is always positive, the
1683: quantity $(r^2 \omega_R -Bm)$ must be negative somewhere for the
1684: integral to vanish. Since $r^2 \omega_R$ increases with $r$, if this
1685: quantity is somewhere negative, then it certainly is negative at the
1686: horizon. In this way we get an upper bound for $\omega_R$:
1687: %
1688: \be
1689: \omega_R<\frac{Bm}{r_H^2}\,.
1690: \label{eq4}
1691: \ee
1692: %
1693: To get a similar bound on $\omega_I$ consider now the real part of
1694: Eq. (\ref{eq2}):
1695: %
1696: \be
1697: \int_{r_H} ^{\infty}dr \left[ \omega_I ^2-\omega_R^2+\frac{2Bm\omega_R}{r^2}
1698: +\frac{c^2fm^2}{r^2}-\frac{B^2m^2}{r^4}+
1699: f^2c^2\frac{|R'| ^2}{|R| ^2}
1700: \right ]\frac{r |R| ^2}{fc^2}=0\,.
1701: \label{eq5}
1702: \ee
1703: %
1704: The positive part of the integrand is
1705: (assuming without loss of generality that $Bm\omega_R>0$)
1706: %
1707: \be
1708: \omega_I ^2+\frac{2Bm\omega_R}{r^2}
1709: +\frac{c^2fm^2}{r^2}+
1710: f^2c^2\frac{|R'| ^2}{|R| ^2}>\omega_I ^2\,.
1711: \label{eq6}
1712: \ee
1713: The negative part is maximized at the horizon:
1714: \be
1715: \omega_R^2+\frac{B^2m^2}{r^4}<\omega_R^2+\frac{B^2m^2}{r_H^4}\,.
1716: \label{eq7}
1717: \ee
1718: %
1719: A necessary condition for the integral to vanish is that the integrand
1720: be negative at the horizon, and this implies
1721: %
1722: \be
1723: \omega_I^2<\omega_R^2+\frac{B^2m^2}{r_H^4}\,,
1724: \label{eq8}
1725: \ee
1726: %
1727: Using the bound (\ref{eq4}) for $\omega_R$ we finally get an upper
1728: bound for $\omega_I$:
1729: %
1730: \be
1731: \omega_I^2<\frac{2B^2m^2}{r_H^4}\,.
1732: \label{eq9}
1733: \ee
1734: %
1735: Notice that the upper bound for $\omega_R$ ensures we are in the
1736: superradiant (or better, superresonant) regime. This seems to be a
1737: general feature, since it has been shown to hold in several black hole
1738: spacetimes. If the instability sets in at all, the system behaves as a
1739: black hole bomb.
1740:
1741: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1742: \begin{thebibliography}{99}
1743:
1744: \bibitem{hawking} S. W. Hawking, Nature {\bf 248}, 30 (1974);
1745: S. W. Hawking, Commun. Math. Phys. {\bf 43}, 199 (1975).
1746:
1747: \bibitem{visserlaws} M. Visser,
1748: Phys. Rev. Lett. {\bf 80}, 3436 (1998).
1749:
1750: \bibitem{visserhawking} M. Visser,
1751: Int. J. Mod. Phys. D {\bf 12}, 649 (2003).
1752:
1753: \bibitem{giddings} S. B. Giddings, S. Thomas, Phys. Rev. D {\bf 65},
1754: 056010 (2002).
1755:
1756: \bibitem{unruh} W. G. Unruh, Phys. Rev. Lett. {\bf 46}, 1351 (1981).
1757:
1758: \bibitem{novello} M. Novello, M. Visser and G. Volovik (editors), {\it
1759: Artificial black holes} (World Scientific, Singapore, 2002).
1760:
1761: \bibitem{visser} M. Visser, Class. Quantum Grav. {\bf 15}, 1767 (1998).
1762:
1763: \bibitem{barcelo} C. Barcel\'o, S. Liberati and M. Visser,
1764: Int. J. Mod. Phys. {\bf A} 18, 3735 (2003).
1765:
1766: \bibitem{unruh2} R. Schutzhold and W. G. Unruh, quant-ph/0408145.
1767:
1768: \bibitem{corleyjacobson} S. Corley and T. Jacobson, Phys. Rev. D {\bf
1769: 59}, 124011 (1999).
1770:
1771: \bibitem{schwinger}
1772: S. Detweiler, Phys. Rev. D {\bf 22}, 2323 (1980); T. J. Zouros and
1773: D. M. Eardley, Annals Phys. {\bf 118}, 139 (1979); T. Damour,
1774: N. Deruelle and R. Ruffini, Lett. Nuovo Cim. {\bf 15}, 257 (1976).
1775:
1776: \bibitem{hod} S. Hod,
1777: Phys. Rev. Lett. {\bf 81}, 4293 (1998).
1778:
1779: \bibitem{kokkotas} K. D. Kokkotas and B. G. Schmidt,
1780: Living Rev. Rel. {\bf2}, 2 (1999);
1781: H.-P. Nollert,
1782: Class. Quantum Grav. {\bf 16}, R159 (1999).
1783:
1784: \bibitem{echeverria} F. Echeverria,
1785: Phys. Rev. D {\bf 40}, 3194 (1989);
1786: L. S. Finn,
1787: Phys. Rev. D {\bf 46}, 5236 (1992).
1788:
1789: \bibitem{price1}
1790: R. H. Price,
1791: Phys. Rev. D{\bf 5}, 2419 (1972).
1792:
1793: \bibitem{schutzhold}
1794: R. Sch\"utzhold, W. G. Unruh,
1795: Phys. Rev. D {\bf 66}, 044019 (2002).
1796:
1797: \bibitem{basak1} S. Basak and P. Majumdar,
1798: Class. Quant. Grav. {\bf 20}, 2929 (2003).
1799:
1800: \bibitem{basak2} S. Basak and P. Majumdar,
1801: Class. Quant. Grav. {\bf 20}, 3907 (2003).
1802:
1803: \bibitem{pressteu} W. Press and S. Teukolsky,
1804: Astrophys. J. {\bf 193}, 443 (1974).
1805:
1806: \bibitem{bhb} V. Cardoso, O. J. C. Dias, J. P.S. Lemos and S. Yoshida,
1807: Phys. Rev. D{\bf 70}, 044039 (2004); {\bf 70}, 044039(E)(2004).
1808:
1809: \bibitem{putten} For a possible astrophysical relevance of the
1810: black hole bomb mechanism, see
1811: M. H. P. M. Putten,
1812: Science {\bf 284}, 115 (1999);
1813: A. Aguirre,
1814: Astrophys. J. {\bf 529}, L9 (2000).
1815:
1816: \bibitem{PGL} P. Painlev\'e, C. R. Hebd. Seances Acad. Sci. {\bf 173},
1817: 677 (1921); A. Gullstrand, Ark. Mat. Astron. Fys. {\bf 16}, 1 (1922);
1818: G. Lema\^itre, Ann. Soc. Sci. Bruxelles, Ser. 1 {\bf 53}, 51 (1933).
1819:
1820: \bibitem{chandra} S. Chandrasekhar, {\it Hydrodynamic and
1821: Hydromagnetic Stability} (Dover Publications, New York, 1981).
1822:
1823: \bibitem{drazin} P. G. Drazin and W. H. Reid, {\it Hydrodynamic
1824: Stability} (Cambridge University Press, 2004).
1825:
1826: \bibitem{willwkb} B. F. Schutz and C. M. Will,
1827: Astrophys. Journal {\bf 291}, L33 (1985);
1828:
1829: \bibitem{will} C. M. Will and S. Iyer,
1830: Phys. Rev. D {\bf 35}, 3621 (1987);
1831: S. Iyer,
1832: Phys. Rev. D {\bf 35}, 3632 (1987);
1833: E. Seidel and Sai Iyer,
1834: Phys.Rev. D {\bf 41}, 374 (1990).
1835:
1836: \bibitem{kokkotaswkb} K. D. Kokkotas,
1837: Class. Quant. Grav. {\bf 8}, 2217 (1991).
1838:
1839: \bibitem{konoplyawkb} R. A. Konoplya,
1840: Phys. Rev. D {\bf 68}, 024018 (2003).
1841:
1842: \bibitem{ferrari} V. Ferrari and B. Mashhoon,
1843: Phys. Rev. {\bf D30}, 295 (1984).
1844:
1845: \bibitem{dreyerfinn} O. Dreyer, B. Kelly, B. Krishnan, L. S. Finn,
1846: D. Garrison and R. Lopez-Aleman, Class. Quantum Grav. {\bf 21}, 787
1847: (2004).
1848:
1849: \bibitem{cardoso4} V. Cardoso, J. P. S. Lemos and S. Yoshida,
1850: gr-qc/0410107.
1851:
1852: \bibitem{dreyer} O. Dreyer,
1853: Phys. Rev. Lett. {\bf 90}, 081301 (2003).
1854:
1855: \bibitem{motl1} L. Motl,
1856: Adv. Theor. Math. Phys. {\bf 6}, 1135 (2003).
1857:
1858: \bibitem{motl2} L. Motl and A. Neitzke,
1859: Adv. Theor. Math. Phys. {\bf 7}, 2 (2003).
1860:
1861: \bibitem{num1} H.-P. Nollert,
1862: Phys. Rev. D {\bf 47}, 5253 (1993);
1863: E. Berti and K. D. Kokkotas,
1864: Phys. Rev. D {\bf 68}, 044027 (2003).
1865:
1866: \bibitem{num3} V. Cardoso, J. P. S. Lemos and S. Yoshida
1867: Phys. Rev. D {\bf 69}, 044004 (2004).
1868:
1869: \bibitem{nils} N. Andersson and C. J. Howls,
1870: Class. Quantum Grav. {\bf 21}, 1623 (2004).
1871:
1872: \bibitem{ricardo} V. Cardoso, J. Nat\'ario and R. Schiappa,
1873: J. Math. Phys. (in press); hep-th/0403132.
1874:
1875: \bibitem{tamaki} T. Tamaki and H. Nomura,
1876: Phys. Rev. D {\bf 70}, 044041 (2004);
1877: J. Kettner, G. Kunstatter and A.J.M. Medved;
1878: gr-qc/0408042;
1879: S. Chen and J. Jing, gr-qc/0409013.
1880:
1881: \bibitem{num2} V. Cardoso, R. Konoplya and J. P. S. Lemos,
1882: Phys. Rev. D {\bf 68}, 044024 (2003);
1883: S. Yoshida and T. Futamase,
1884: Phys. Rev. D {\bf 69}, 064025 (2004).
1885: R. A. Konoplya and A. Zhidenko,
1886: JHEP {\bf 0406}, 037 (2004).
1887:
1888: \bibitem{neitzke} We thank Andrew Neitzke for sharing with us a
1889: related unpublished calculation for the Kerr metric.
1890:
1891: \bibitem{TedEinsteinEOS} T. Jacobson, Phys. Rev. Lett. {\bf 75}, 1260
1892: (1995).
1893:
1894: \bibitem{BLSV} C. Barcel\'o, S. Liberati, S. Sonego and M. Visser,
1895: gr-qc/0408022.
1896:
1897: \bibitem{price2}
1898: C. Gundlach, R. H. Price and J. Pullin,
1899: Phys. Rev. D{\bf 49}, 883 (1994);
1900: C. Gundlach, R. H. Price and J. Pullin,
1901: Phys. Rev. D{\bf 49}, 890 (1994);
1902: S. Hod,
1903: Phys. Rev. D{\bf 58}, 104022 (1998);
1904: L. Barack and A. Ori,
1905: Phys. Rev. Lett. {\bf 82}, 4388 (1999);
1906: N. Andersson and K. Glampedakis,
1907: Phys. Rev. Lett. {\bf 84}, 4537 (2000);
1908: H. Koyama and A. Tomimatsu,
1909: Phys. Rev. D{\bf 64}, 044014 (2001);
1910: S. Hod,
1911: Class. Quant. Grav. {\bf 18}, 1311 (2001).
1912:
1913: \bibitem{ching1} E. S. C. Ching, P. T. Leung, W. M. Suen and K. Young,
1914: Phys. Rev. Lett. {\bf 74}, 2414 (1995).
1915:
1916: \bibitem{ching2} E. S. C. Ching, P. T. Leung, W. M. Suen and K. Young,
1917: Phys. Rev. D{\bf 52}, 2118 (1995).
1918:
1919: \bibitem{cardosoDdimtails} V. Cardoso, S. Yoshida, O. J. C. Dias and
1920: J. P.S. Lemos, Phys. Rev. D {\bf 68}, 061503 (2003).
1921:
1922: \bibitem{ginzburg} V. L. Ginzburg and I. M. Frank, Dokl. Akad. Nauk
1923: SSSR {\bf 56}, 583 (1947); for a recent review cf. V. L. Ginzburg, in
1924: {\it Progress in Optics XXXII}, edited by E. Wolf (Elsevier,
1925: Amsterdam, 1993).
1926:
1927: \bibitem{bekschiffer}
1928: J. D. Bekenstein and M. Schiffer,
1929: Phys. Rev. D {\bf 58}, 064014 (1998).
1930:
1931: \bibitem{zeldovich}
1932: Ya. B. Zel'dovich, JETP Lett. {\bf 14}, 180 (1971); Sov. Phys. JETP
1933: {\bf 35}, 1085 (1972).
1934:
1935: \bibitem{superr}
1936: C. W. Misner,
1937: Phys. Rev. Lett. {\bf 28}, 994 (1972);
1938: A. A. Starobinsky,
1939: Sov. Phys. JETP {\bf 37}, 28 (1973);
1940: A. A. Starobinsky and S. M. Churilov,
1941: Sov. Phys. JETP {\bf 38}, 1 (1973);
1942: W. Unruh,
1943: Phys. Rev. D {\bf 10}, 3194 (1974);
1944: W. H. Press and S. A. Teukolsky,
1945: Astrophys. Journal {\bf 185}, 649 (1973).
1946:
1947: \bibitem{beksuperr}
1948: J. D. Bekenstein,
1949: Phys. Rev. D {\bf 7}, 949 (1973).
1950:
1951: \bibitem{penrose}
1952: R. Penrose, Nuovo Cimento {\bf 1}, 252 (1969).
1953:
1954: \bibitem{friedman}
1955: J. L. Friedman,
1956: Commun. Math. Phys. {\bf 63}, 243 (1978).
1957:
1958: \bibitem{cominsschutz}
1959: N. Comins and B. F. Schutz,
1960: Proc. R. Soc. Lond. A {\bf 364}, 211 (1978).
1961:
1962: \bibitem{ptbhb}
1963: W. H. Press and S. A. Teukolsky,
1964: Nature {\bf 238}, 211 (1972).
1965:
1966: \bibitem{tedvolovik}
1967: T. Jacobson and G. E. Volovik,
1968: Phys. Rev. D {\bf 58}, 064021 (1998).
1969:
1970: \bibitem{alp}
1971: N. Andersson, P. Laguna and P. Papadopoulos, Phys. Rev. D {\bf 58},
1972: 087503 (1998).
1973:
1974: \bibitem{abramowitz}
1975: M. Abramowitz and A. Stegun, {\it Handbook of mathematical functions}
1976: (Dover Publications, New York, 1970).
1977:
1978: \bibitem{det} S. L. Detweiler and J. R. Ipser,
1979: Astrophys. Journal {\bf 185}, 675 (1973).
1980: \end{thebibliography}
1981:
1982: \end{document}
1983: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1984:
1985: