gr-qc0505149/BBH1.tex
1: %\documentclass[aps,prd, galley, showpacs, amssymb]{revtex4}
2: %\documentclass[aps,prd, preprint, showpacs, amssymb]{revtex4}
3: \documentclass[aps,prd, twocolumn, showpacs, amssymb, floatfix]{revtex4}
4: \usepackage{graphicx,amsmath,subfigure}
5: \begin{document}
6: 
7: \author{Ilya Mandel}
8: \affiliation{Theoretical Astrophysics, 
9: 	California Institute of Technology, Pasadena, CA 91125}
10: \email{ilya@caltech.edu}
11: \date{May 28, 2005}
12: \title{The geometry of a naked singularity created by standing waves
13: near a Schwarzschild horizon, and its application to the binary black hole 
14: problem}
15:  
16: \begin{abstract}
17: 
18: The most promising way to compute the gravitational waves emitted by binary 
19: black holes (BBHs) in their last dozen orbits, where post-Newtonian techniques 
20: fail, is a {\it quasistationary approximation} introduced by Detweiler and 
21: being pursued by Price and others.  In this approximation the outgoing
22: gravitational waves at infinity and downgoing gravitational waves at the holes'
23: horizons are replaced by standing waves so as to guarantee that the
24: spacetime has a helical Killing vector field.  Because the horizon generators
25: will not, in general, be tidally locked to the holes' orbital motion, the 
26: standing waves will destroy the horizons, converting the black holes into 
27: naked singularities that resemble black holes down to near the horizon radius.
28: This paper uses a spherically symmetric, scalar-field model problem
29: to explore in detail the following BBH issues: (i) The destruction of a horizon 
30: by the standing waves. (ii) The accuracy with which the resulting naked 
31: singularity resembles a black hole. (iii) The conversion of the standing-wave
32: spacetime (with a destroyed horizon) into a spacetime with downgoing waves 
33: by the addition of a ``radiation-reaction field''. (iv) The accuracy with 
34: which the resulting downgoing waves agree with the downgoing waves of a 
35: true black-hole spacetime (with horizon).  The model problem used to study 
36: these issues consists of a Schwarzschild black hole endowed with spherical 
37: standing waves of a scalar field, whose wave frequency and near-horizon 
38: energy density are chosen to match those of the standing gravitational 
39: waves of the BBH quasistationary approximation.   It is found
40: that the spacetime metric of the singular, standing-wave spacetime, and 
41: its radiation-reaction-field-constructed downgoing waves are quite close to
42: those for a Schwarzschild black hole with downgoing waves --- sufficiently 
43: close to make the BBH quasistationary approximation look promising for
44: non-tidally-locked black holes.
45: 
46: \end{abstract}
47: 
48: \pacs{04.25.-g, 04.25.Nx, 04.30.-w}
49: 
50: \maketitle
51: 
52: \section{\label{sec:intro}Introduction and summary}
53: 
54: It is very important, in gravitational astronomy, to have accurate
55: computations of the gravitational waves from the inspiral of a black 
56: hole binary \cite{GravRad}.  However, computing these waves is extremely
57: challenging: for the last $\approx 25$ cycles of inspiral waves, post-Newtonian
58: approximations fail \cite{Brady}, and numerical relativity can not yet 
59: evolve the full dynamical equations in this regime.  It appears that the 
60: best hope for accurately computing the BBH inspiral waves is by a
61: quasi-stationary approximation \cite{Detweiler, Price}.  In this approximation,
62: the energy and angular momentum of the binary are conserved by the imposition 
63: of standing gravitational waves, and the spacetime has a helical Killing vector 
64: field.  The standing-wave radiation required to keep the orbit stationary is 
65: computed by demanding that the energy contents of the gravitational waves (GW) 
66: be minimized \cite{Price}.  
67: 
68: Standing-wave radiation consists of a sum of ingoing and outgoing radiation at 
69: infinity, and downgoing and upgoing radiation at the black-hole horizons.  
70: The physical spacetime, with purely outgoing waves at infinity and downgoing 
71: waves at the horizons, can be recovered from the standing-wave spacetime by 
72: adding a perturbative radiation-reaction field \cite{private}.  The solution 
73: for the BBH inspiral consists of a series of quasi-stationary solutions that 
74: evolve from one to another via energy and angular momentum loss triggered by 
75: the radiation-reaction field.  The waves measured at a detector can be deduced 
76: from this sequence of quasistationary solutions.
77: 
78: The black holes comprising the binary are tidally locked if their horizon 
79: generators are static in the frame co-rotating with the orbit.
80: In the tidally locked case, the metric perturbations necessary to keep the 
81: black holes on a stationary orbit are static in the co-rotating frame,
82: and the black holes can be regarded as having bifurcate Killing horizons
83: (both a past horizon and a future horizon).  
84: 
85: In reality, the black holes are not tidally locked.  Their mutual tidal forces 
86: are not strong enough to maintain locking during the inspiral.  In the absence 
87: of tidal locking, the standing waves of the standing-wave approximation 
88: destroy the black-hole horizons: the downgoing waves destroy the past horizon 
89: by building up an infinite energy density at the past horizon, and the 
90: upgoing waves destroy the future horizon.  Therefore, we expect that forcing 
91: the orbit to be stationary via the addition of standing gravitational waves 
92: will strip the Kerr black holes of their horizons and leave naked 
93: singularities in their place \cite{Wald}.
94: 
95: Despite this radical change in the character of the orbiting bodies, it is 
96: reasonable to expect that the standing-wave solution will give a quite
97: accurate approximation to the true physical black-hole spacetimes everywhere
98: except very near the black-hole horizons.  In order to verify or refute this
99: expectation, it is necessary to explore the nature of the singularities 
100: created by the standing gravitational waves and to test how well the physical 
101: solution with true black holes can be extracted from the standing-wave 
102: solution with naked singularities.  
103: 
104: As a first step in such an exploration, we consider in this paper a simple 
105: model problem designed to give insight into the nature of the singularities 
106: generated by the standing gravitational waves, and the accuracy with which 
107: the physical, BBH spacetime can be recovered from the standing-wave, 
108: singularity-endowed spacetime.  
109: 
110: Our model problem is a single, spherically symmetric black hole that is 
111: converted into a naked singularity by spherical standing waves of a scalar 
112: field. 
113: 
114: We begin our analysis in Sec.~\ref{sec:sfeq} by describing the mapping 
115: between the BBH problem, into which we seek insight, and our spherical, 
116: scalar-field model problem. In particular, we deduce what should be the 
117: range of scalar-field amplitudes and frequencies in order to mock up the 
118: gravitational waves of the BBH problem.
119: 
120: Then in Sec.~\ref{sec:sw}, we construct and explore the standing-wave 
121: spacetime for our spherical model problem.  We initially treat the 
122: standing-wave scalar field as residing in the unperturbed Schwarzschild
123: spacetime of the black hole, and we use Regge-Wheeler first-order perturbation 
124: theory to compute the scalar-energy-induced deviations of the hole's metric
125: from Schwarzschild.  The metric perturbations consist of a static component 
126: and a component varying in time at twice the scalar-field frequency 
127: (see Fig.~\ref{Hpert-fig} below).  The oscillatory component is smaller 
128: than the static one and higher-order harmonics of both the field and 
129: the metric are strongly suppressed.  
130: 
131: The static metric perturbation grows divergently as one approaches 
132: the Schwarzschild horizon --- an obvious indication of the horizon's 
133: destruction by the standing-wave stress-energy.  To explore the structure 
134: of the resulting naked singularity, in Sec.~\ref{sec:sw-nonlin} we abandon 
135: perturbation theory and switch to the fully nonlinear, coupled Einstein 
136: equations and scalar-field equations.  To simplify the analysis, we focus 
137: solely on the static part of the singularity's metric; we do this by time 
138: averaging the scalar stress-energy tensor before inserting it into the 
139: fully nonlinear Einstein equations.  We solve the resulting equations 
140: numerically to obtain the spacetime geometry outside the singularity.
141: The geometry's embedding diagram (Fig.~\ref{emb-fig} below) and the redshift 
142: seen by a distant observer (Fig.~\ref{red-fig} below) show that the spacetime 
143: remains nearly Schwarzschild outside the Schwarzschild horizon, but deviates 
144: strongly from Schwarzschild at $r\approx 2M$ and below.  (Here $M$ is the mass 
145: of the hole-like singularity and we use geometrized
146: units $c=G=1$ everywhere in this paper.)  Above $r = 2M$, the standing-wave 
147: spacetime is very nearly identical to the Schwarzschild spacetime down to 
148: radii that are well inside the inner edge of the effective potential 
149: (Fig.~\ref{spacetime-fig}). Below $r = 2M$, radial distance 
150: changes far more slowly than areal radius; i.e., $g_{rr}$ tends to $0$ as 
151: $r \rightarrow 0$.  The redshift seen by an external observer rises rapidly 
152: when the emitter falls inside $r = 2M$.  However, a signal from the 
153: singularity at $r=0$ may be infinitely redshifted or infinitely blueshifted, 
154: depending on the choice of scalar field parameters.
155: 
156: In Sec.~\ref{sec:down} we turn to the model spherical spacetime 
157: that mocks up our desired BBH solution: the spacetime of a Schwarzschild
158: black hole with downgoing scalar waves.  Not surprisingly, the metric 
159: perturbations induced by the downgoing scalar-wave energy are those of the 
160: Vaidya solution of Einstein's equations --- a slowly growing black hole
161: with a smooth, non-singular future horizon.  This spacetime is well 
162: approximated, for short time intervals, by the Schwarzschild solution 
163: with (constant) Schwarzschild mass equal to the instantaneous Vaidya mass.
164: 
165: Finally, in Sec.~\ref{sec:reconstruct} we demonstrate that by adding a 
166: perturbative radiation-reaction field to the standing-wave solution, a  
167: downgoing solution to the scalar-wave equation can be recovered.  
168: We explore the level of agreement between these downgoing waves that live in 
169: the singularity-endowed standing-wave spacetime and the downgoing
170: waves in the Schwarzschild approximation to the Vaidya spacetime. 
171: The agreement (for details see Sec.~\ref{sec:reconstruct} 
172: and Fig.~\ref{recon-fig} below) is rather good for scalar-wave amplitudes 
173: and frequencies that mock up the BBH problem --- sufficiently good to 
174: give optimism that the standing-wave approximation will give accurate 
175: gravitational waveforms for the final stages of binary-black-hole inspiral.
176: 
177: 
178: 
179: \section{\label{sec:sfeq}
180: The mapping between the BBH problem and our model scalar-field problem}
181: 
182: In our exploration of the quasistationary, standing-wave approximation for
183: black-hole binaries we shall study several spherically symmetric 
184: spacetimes, each endowed with a standing-wave scalar field.  In 
185: Sec.~\ref{sec:sw-pert} the spacetime will be Schwarzschild, or Schwarzschild
186: with first-order gravitational perturbations generated by the scalar-field
187: stress-energy tensor.  In Sec.~\ref{sec:sw-nonlin} the spacetime will be
188: that of a naked singularity generated by the coupled, time-averaged 
189: Einstein-scalar-field equations.  In this section we shall identify the 
190: parameter regime relevant to gaining insight from these spacetimes into
191: the binary black hole problem.  
192: 
193: In each of these spherical spacetimes, the scalar field must be a 
194: solution to the wave equation:
195: \begin{equation}\label{wave}
196: \Box \Phi =
197: \frac{1}{\sqrt{-g}}(\sqrt{-g} g^{\alpha\beta} \Phi_{,\alpha})_{,\beta}=
198: 0 \, ,
199: \end{equation}
200: where $g_{\alpha\beta}$ is the spacetime metric with the interval
201: \begin{equation}\label{genmetric}
202: ds^2=f(r,t) dt^2 +g(r,t) dr^2 +r^2 (d\theta^2 +\sin^2 \theta d\phi^2) .
203: \end{equation}
204: We assume that the scalar field is monochromatic with frequency $\omega$,
205: and we write it in the form
206: \begin{equation}\label{Phi}
207: \Phi=\Re\left ( \frac{\Psi(r) e^{-i\omega t}}{r} \right) \, ,
208: \end{equation}
209: where $\Re()$ denotes the real part and the phase was set by the choice
210: of the zero of time $t$. 
211: 
212: The scalar field $\Phi$ serves as the source of curvature in the Einstein 
213: equations,
214: \begin{equation}\label{Einstein}
215: G_{\alpha\beta}=8\pi T_{\alpha\beta} \, ,
216: \end{equation}
217: where the stress-energy tensor depends on the scalar field according to
218: \begin{equation}\label{T}
219: T_{\alpha\beta}=\frac{1}{4\pi} \Phi_{,\alpha} \Phi_{,\beta}-
220: 	\frac{1}{8\pi}g_{\alpha\beta} \Phi_{,\mu} \Phi^{,\mu}
221: \end{equation}
222: (cf.~Eq.~(20.66) of \cite{MTW} or Eq.~(A.11) of \cite{NF}).
223: 
224: We can re-write equations (\ref{Einstein}) and (\ref{T}) in a simpler form via 
225: the Ricci tensor:
226: \begin{equation}\label{Ricci}
227: R_{\alpha\beta}=2 \Phi_{,\alpha} \Phi_{,\beta} \, .
228: \end{equation}
229: 
230: Relevant ranges for the scalar-field frequency and amplitude are determined 
231: by the binary black hole problem we are modeling.  Suppose that the black 
232: holes in the binary have equal mass $M$, and let $a$ be their radial 
233: separation.  Since we are interested in the late inspiral, where the 
234: post-Newtonian  methods fail, the desired range of parameters should 
235: correspond to $6 \alt a/M \alt 15$ \cite{Brady}.  
236: 
237: The Keplerian orbital frequency of the black holes is
238: \begin{equation}\label{omegaK} 
239: \Omega=\frac{1}{M} \sqrt{\frac{2}{(a/M)^3}} \, .
240: \end{equation}
241: The gravitational wave frequency is twice the Keplerian frequency, and we 
242: set our scalar-field frequency equal to the GW frequency:   
243: \begin{equation}\label{omega}
244: \omega = 2\Omega = \frac{2}{M} \sqrt{\frac{2}{(a/M)^3}} \, .
245: \end{equation}
246: 
247: The power going down a black hole due to the orbital motion of its companion 
248: is approximately
249: \begin{equation}\label{power}
250: P_{GW} = \frac{32}{5} M^4 \mu^2 \Omega^6,
251: \end{equation}
252: where $\mu$ is the mass of the companion \cite{Poisson, Hua}.  Although the 
253: calculations in Refs.~\cite{Poisson, Hua} underlying Eq.~(\ref{power}) were 
254: carried out under the assumption  $\mu \ll M$, we will use Eq.~(\ref{power}) to 
255: approximate the power for equal mass black holes, $\mu=M$.  This approximation 
256: is not too worrisome because we are interested in the general features of the 
257: scalar-field model, which roughly corresponds to the interesting range of BBH 
258: separations, rather than in the quantitative results for this model.  We 
259: select the scalar-field amplitude by demanding that its energy density near 
260: the horizon equal the GW energy density there:
261: \begin{equation}\label{dEdV}
262: \frac{dE}{dV} \approx \frac{P_{GW}}{4\pi(2M)^2} \, .
263: \end{equation}
264: (In the spirit of this approximate analysis we here ignore the gravitational 
265: blueshift of the energy.)   By equating this energy density to the value of
266: $T_{00}$ at the horizon, computed by inserting Eq.~(\ref{Phi}) into 
267: Eq.~(\ref{T}), we obtain the scalar-field amplitude inside 
268: the peak of the effective potential:
269: \begin{equation}\label{amp}
270: \Psi_{in} = \sqrt{\frac {64}{5}} \left[\frac{1}{(a/M)}\right]^3 M \, .
271: \end{equation}
272: 
273: Using equations (\ref{omega}) and (\ref{amp}), we can compute the desired
274: scalar-field frequency and amplitude for the boundaries of the
275: region of interest:
276: \begin{subequations}\label{arange}
277: \begin{equation}\label{a6}
278: a = 6 M  \Rightarrow \omega \approx 0.19/M, \, \Psi_{in} \approx 0.017 M ;
279: \end{equation}
280: \begin{equation}\label{a15}
281: a = 15 M  \Rightarrow \omega \approx 0.049/M, \, \Psi_{in} \approx 0.0011 M .
282: \end{equation}
283: \end{subequations}
284: 
285: 
286: \section{\label{sec:sw}Standing-wave scalar field}
287: 
288: We now turn to the standing-wave scalar-field spacetime that mocks up the
289: spacetimes of the BBH standing-wave approximation.  The metric of this 
290: spacetime has the form of Eq.~(\ref{genmetric}) and the standing-wave 
291: scalar field follows from Eq.~(\ref{Phi}):
292: \begin{equation}\label{Phistat}
293: \Phi=\frac{\Psi(r) \cos{\omega t}}{r} \, ,
294: \end{equation}
295: where $\Psi(r)$ is now real.  
296: 
297: We shall treat the standing-wave scalar field twice via two different 
298: simplifying  assumptions.  First, in Sec.~\ref{sec:sw-pert}, we will consider 
299: the scalar field perturbatively; its wave equation will be that of 
300: the Schwarzschild spacetime, and its stress-energy will generate first-order
301: perturbations of the metric away from Schwarzschild.  Then in 
302: Sec.~\ref{sec:sw-nonlin}, we will consider the fully nonlinear
303: Einstein-scalar-field spacetime but with the scalar stress-energy averaged over
304: time to make the metric static.
305: 
306: \subsection{\label{sec:sw-pert}Perturbative standing-wave solution}
307: 
308: \subsubsection{Perturbative formalism for the standing-wave spacetime}
309: 
310: In our first approach, the lowest-order solution for the scalar field is 
311: computed by solving the wave equation (\ref{wave}) in the Schwarzschild 
312: background with the metric
313: \begin{eqnarray}\label{metricB}
314: ds^2 &=& g_{\alpha\beta}^B dx^\alpha dx^\beta \\
315: \nonumber
316: 	&=& -(1-2/r)dt^2+\frac{1}{1-2/r}dr^2+r^2d\Omega^2 \, ,
317: \end{eqnarray}
318: where we rescale so that $M=1$.
319: The wave equation simplifies as follows (cf.~Eq.~(32.27b) of \cite{MTW}):
320: \begin{equation}\label{wavepert}
321: \frac{d^2 \Psi}{{d r^*}^2}=\left[-\omega^2+(1-2/r)\frac{2}{r^3}\right] \Psi \, ,
322: \end{equation}
323: where $r^*$ is the Regge-Wheeler tortoise coordinate \cite{RW}, 
324: \begin{equation}\label{rRW}
325: r^*=r+2 \ln{(r/2-1)} \, .
326: \end{equation}
327: Because $\omega^2$ dominates the right hand side of Eq.~(\ref{wavepert}) 
328: both far from the horizon ($r \gg 2$) and very near the horizon, the 
329: scalar field will oscillate with a nearly constant frequency $\omega$ in 
330: those regions.  In between, where the effective potential 
331: \begin{equation}\label{V}
332: V(r^*)=(1-2/r)(2/r^3) ,
333: \end{equation} 
334: is significant, there is an intermediate transitional region 
335: (see Fig.~\ref{Psi-fig}).  (In this paper we mention several times ``the inner
336: edge of the peak of the effective potential''; we define this inner edge to be
337: the radius at which the effective potential drops to one percent of 
338: its maximum value at the peak.)
339: 
340: \begin{figure}[ht]
341: \includegraphics[keepaspectratio=true,width=8.5cm,angle=0]{wavepert.eps}
342: \caption{The standing-wave scalar field in a Schwarzschild background (solid
343: curve) and the effective potential (dashed curve) for angular frequency 
344: $\omega=0.049$.}
345: \label{Psi-fig}
346: \end{figure}
347: 
348: Since we are approaching the problem perturbatively, we are interested in 
349: some  small metric perturbation $h_{\alpha\beta}$ on top of the 
350: background metric $g^B_{\alpha\beta}$ of Eq.~(\ref{metricB}) that would 
351: yield the curvature corresponding  to the stress-energy tensor of the scalar 
352: field:  
353: \begin{equation}\label{metricpert}
354: g_{\alpha\beta}=g^B_{\alpha\beta} + h_{\alpha\beta} \, .
355: \end{equation}
356: Linearizing in $h_{\alpha\beta}$, this metric gives the Ricci tensor
357: %(cf. Eq.~(5.14) of \cite{GW})
358: \begin{equation}\label{Riccipert}
359: R_{\alpha\beta}=R^B_{\alpha\beta} + \frac{1}{2} \left(
360: 	h_{\mu\alpha | \beta}^{\phantom{\mu\alpha | \beta}\mu} + 
361: 	h_{\mu\beta| \alpha}^{\phantom{\mu\beta| \alpha}\mu}
362: 	- h_{\alpha\beta | \mu}^{\phantom{\alpha\beta | \mu}\mu} 
363: 	- h_{| \alpha\beta} \right) \, ,
364: \end{equation}
365: where $h=h_{\mu}^{\phantom{\mu}\mu}$ and $_{|}$ represents the covariant 
366: derivative in the background metric $g^B_{\alpha\beta}$.
367: For the Schwarzschild background metric, $R^B_{\alpha\beta}=0$.
368: 
369: We are interested only in spherically symmetric perturbations.
370: A gauge transformation brings additional simplification, so 
371: $h_{\alpha\beta}$ can be written in the following simple 
372: Regge-Wheeler form:
373: \begin{equation}\label{hgennoH1K}
374: h_{\alpha\beta}= \left( \begin{array}{*{4}{c}} (1-2/r) H_0(t,r) & 0 & 0 & 0 \\
375: 					0 & \frac{H_2(t,r)}{1-2/r} & 0 & 0 \\
376: 					0 & 0 & 0 & 0 \\
377: 					0 & 0 & 0 & 0 
378: 			\end{array}  \right) \, .
379: \end{equation}
380: (Compare with Eq.~(13) of \cite{RW} for the case $L=0$.)
381: 
382: We can now substitute $h_{\alpha\beta}$ given by Eq.~(\ref{hgennoH1K}) into 
383: Eq.~(\ref{Riccipert}) to compute the perturbed Ricci tensor:
384: \begin{subequations}\label{RiccinoH1K}
385: \begin{eqnarray}
386: R_{tt}&=&-\frac{1}{2}\frac{\partial^2}{\partial t^2} H_2 
387: \\
388: \nonumber
389: \lefteqn{
390: -\frac{r-2}{2 r^3}
391: 	\left[ (2r-1)\frac{\partial}{\partial r} H_0 +
392: 		\frac{\partial}{\partial r}H_2 +
393: 		r(r-2)\frac{\partial^2}{\partial r^2}H_0 \right] \, ;
394: }
395: \\
396: R_{tr}&=&\frac{1}{r} \frac{\partial}{\partial t}H_2 \, ;
397: \\
398: R_{rr}&=&\frac{r^2}{2(r-2)^2} \frac{\partial^2}{\partial t^2} H_2 + 
399: 	\frac{1}{2r(r-2)}
400: \\
401: \nonumber
402: \lefteqn{\times
403: 	\left[ 3\frac{\partial}{\partial r} H_0 +
404: 	(2r-3)\frac{\partial}{\partial r} H_2 +
405: 	r(r-2)\frac{\partial^2}{\partial r^2} H_0 \right] \, ;	
406: }
407: \\
408: R_{\theta\theta}&=&H_2 +\frac{r-2}{2} \frac{\partial}{\partial r} H_0 + 
409: 	\frac{r-2}{2} \frac{\partial}{\partial r}H_2 \, .\qquad \qquad \qquad
410: \end{eqnarray}
411: \end{subequations}
412: 
413: Inserting expressions (\ref{RiccinoH1K}) for $R_{\alpha\beta}$ into the 
414: Einstein equations (\ref{Ricci}), one obtains a set of rather complicated 
415: PDE's containing both spatial and time derivatives to the second order.  
416: However, we expect that the equations can be further simplified because of 
417: additional consistency conditions imposed on $\Phi$ by the wave equation 
418: (\ref{wavepert}).  Indeed, after adding the $R_{tt}$ and $R_{rr}$ equations
419: with appropriate coefficients to remove the second derivatives in both
420: $t$ and $r$, and using $R_{\theta\theta}=0$ to relate $H_0$ to
421: $H_2$, we obtain the following set of first-order ODE's for $H_0$ and $H_2$:
422: \begin{subequations}\label{H0H2}
423: \begin{equation}\label{H2}
424: \frac{\partial H_2}{\partial r} = -\frac{H_2}{r-2} + 
425: 	\frac{r^3}{(r-2)^2} \Phi_{, t} \Phi_{,t} +
426: 	r \Phi_{, r} \Phi_{, r} \, ;
427: \end{equation}
428: \begin{equation}\label{H0}
429: \frac{\partial H_0}{\partial r} = -\frac{\partial H_2}{\partial r} 
430: 				-\frac{2}{r-2}H_2 \, .
431: \end{equation}
432: \end{subequations}
433: These far simpler equations can be shown to produce no spurious solutions; 
434: in fact, together with the wave equation (\ref{wavepert}),
435: they are equivalent to the second-order PDE system (\ref{RiccinoH1K}).
436: 
437: \subsubsection{First-order metric perturbations due to the standing-wave 
438: scalar field}
439: In the scalar-field ansatz (\ref{Phistat}) we assumed $\Phi \propto 
440: \cos{\omega t}$.  Therefore, the driving term on the right hand 
441: side of Eq.~(\ref{H2}) will have static components as well as components 
442: oscillating in time at the frequency $2\omega$.  Because there is no 
443: mixing of terms with distinct time signatures in equations (\ref{H0H2}), 
444: these terms may be treated separately:
445: \begin{subequations}\label{Hsplit}
446: \begin{eqnarray}
447: H_2(t,r) &=& H_2^{stat} (r) + H_2^{cos} (r)  \cos{2\omega t} \, ;
448: \\
449: H_0(t,r) &=& H_0^{stat} (r) + H_0^{cos} (r)  \cos{2\omega t} \, .
450: \end{eqnarray}
451: \end{subequations}
452: (There is no $\sin{2\omega t}$ term with our particular choice of the 
453: scalar-field phase.)
454: 
455: For $r \gg 2$ analytical approximations for $H_0$ and $H_2$ are easy to obtain
456: because the scalar field is particularly simple there: 
457: \begin{subequations}\label{Hfar}
458: \begin{equation}
459: \Phi \approx (\Psi_0/r) \cos{(\omega r^*)} \cos{(\omega t)} ,
460: \end{equation}
461: where $\Psi_0$ is the scalar-field amplitude as $r \rightarrow \infty$.
462: Inserting this into Eqs.~(\ref{H0H2}), we readily compute, at large $r$:
463: \begin{eqnarray}
464: H_2^{stat} (r) &\approx& \frac{1}{2}\omega^2 \Psi_0^2 - \frac{\Psi_0^2}{4r^2} 
465: 	- \frac{\Psi_0^2\cos{2\omega r^*}}{4r^2} \, ;\\
466: H_2^{cos} (r) &\approx& -\frac{\Psi_0^2}{4r^2} - 
467: 	\frac{\Psi_0^2\cos{2\omega r^*}}{4r^2} \\
468: 	\nonumber
469: 	&&-\frac{\Psi_0^2 \omega \sin{2\omega r^*}}{4r} \, ;\\
470: H_0^{stat} (r) &\approx& -\omega^2 \Psi_0^2 \ln{r} + 
471: 	\frac{\Psi_0^2\cos{2\omega r^*}}{4r^2}  \, ;\\
472: H_0^{cos} (r) &\approx& \frac{\Psi_0^2\cos{2\omega r^*}}{4r^2} + 
473: 	\frac{\Psi_0^2 \omega \sin{2\omega r^*}}{4r} \, .
474: \end{eqnarray}
475: \end{subequations}
476: The static components of $H_2$ and $H_0$ are non-vanishing at infinity, 
477: and $H_0^{stat}$ actually diverges.  This indicates that, due to the energy
478: contained in the scalar field, the spacetime is not asymptotically flat. 
479: However, this bad behavior at infinity is an artifact of our model problem 
480: and is irrelevant to the issues we are studying in this paper.
481: 
482: We can read off from Eqs.~(\ref{Hfar}) the ratios of the 
483: oscillatory and static components of the metric perturbations at large $r$.  
484: They are 
485: \begin{subequations}
486: \begin{equation}
487: \left|\frac{H_2^{cos}}{H_2^{stat}}\right| \approx \frac{1}{2\omega r}
488: \end{equation}
489: and 
490: \begin{equation}
491: \left|\frac{H_0^{cos}}{H_0^{stat}}\right| \approx \frac{1}{4\omega r \ln{r}}\, ;
492: \end{equation}
493: \end{subequations}
494: thus, the static components dominate far from the horizon.
495: 
496: Equations (\ref{Hfar}) can be used to set initial conditions for the 
497: metric perturbations at some large $r$, allowing for a numerical solution to 
498: Eqs.~(\ref{H0H2}) from there down to the horizon, $r=2$.  The resulting 
499: solution, plotted in  Fig.~\ref{Hpert-fig}, indicates that static components 
500: continue to dominate near the horizon.
501: 
502: \begin{figure}[ht]
503: \includegraphics[keepaspectratio=true,width=8.5cm,angle=0]{metricpert.eps}
504: \caption{Metric perturbations for a standing-wave scalar field in a 
505: 	Schwarzschild background with angular frequency $\omega=0.19$
506: 	and amplitude $\Psi_0=0.015$ far from the black hole,
507: 	corresponding to a binary separation $a \approx 6 M$,
508: 	[Eq.~(\ref{a6})].}
509: \label{Hpert-fig}
510: \end{figure}
511: 
512: Near the horizon (inside the effective-potential peak), the scalar field has
513: the form 
514: \begin{subequations}
515: \begin{equation}
516: \Phi \approx (\Psi_{in}/2) \cos{\omega r^*} \cos{\omega t} ,
517: \end{equation} 
518: where $\Psi_{in}$ is the scalar-field amplitude as $r \rightarrow 2$.
519: Inserting this approximation into Eq.~(\ref{H2}) and averaging the right-hand
520: side leads to the following rough estimate of the magnitude of the 
521: perturbation near the horizon:
522: \begin{equation}\label{H2near}
523: H_2^{stat} \approx \frac{2 \omega^2 \Psi^2 \ln{(r-2)}} {r-2} \, .
524: \end{equation}
525: \end{subequations}
526: Inverting this formula can give a useful estimate of the distance from 
527: the horizon where the perturbation reaches a particular value; 
528: the estimate turns out to be accurate to within a factor of two.
529: 
530: Although it appears that the metric perturbation diverges at the expected 
531: location of the horizon, our perturbative solution is not trustworthy in 
532: this regime for several reasons in addition to the obvious one of
533: violating the perturbative assumption $H_0, H_2 \ll 1$:
534: 
535: 1. We ignored the {\it back reaction}, i.e., the feedback of the metric 
536: perturbation into the wave equation.  Using the Schwarzschild metric in place 
537: of the more accurate perturbed metric in the wave equation, that
538: is, using the approximate Eq.~(\ref{wavepert}) in place of the exact 
539: Eq.~(\ref{wave}), is equivalent to an error in the scalar-field
540: frequency $\Delta \omega / \omega \approx O(H)$, which produces phase offsets 
541: in the scalar field when the wave equation is integrated numerically.  
542: 
543: 2. We linearized the Ricci tensor in the perturbations, neglecting 
544: higher-order $O(H^2)$ effects. In contrast to the linearized equations 
545: (\ref{H0H2}) for $H_2$ and $H_0$, the nonlinear perturbative equations are:
546: \begin{subequations}
547: \begin{eqnarray}
548: \label{H2nonlin}
549: \frac{\partial H_2}{\partial r} &=& -\frac{H_2 (1+H_2)}{r-2}
550: \\
551: \nonumber
552: & &
553: 	+\frac{r^3}{(r-2)^2} \frac{(1+H_2)^2}{1-H_0} (\Phi_{, t})^2 +
554: 	r (1+H_2) (\Phi_{, r})^2 \, ;
555: \\
556: \label{H0nonlin}
557: \frac{\partial H_0}{\partial r} &=& (1-H_0) \left[ 
558: 	-\frac{1}{1+H_2}\frac{\partial H_2}{\partial r} -
559: 	\frac{2 H_2}{r-2} \right] \, .
560: \end{eqnarray}
561: \end{subequations}
562: Linearization introduces local errors of order $H$ into the Einstein 
563: equations.  However, the errors can build up globally when the 
564: equations are integrated to obtain a numerical solution.  The errors produced
565: by linearizing the Ricci tensor (the differences between solutions to the
566: linearized and nonlinear Einstein equations without back reaction in the 
567: wave equation) have the same order of magnitude in the parameter range of 
568: interest as the errors produced by neglecting {\it back reaction} (the 
569: differences between solutions to the nonlinear Einstein equations 
570: depending on whether wave equation (\ref{wavepert}) or (\ref{wave}) is used). 
571: 
572: 3. We ignored higher harmonics of the scalar field and of the metric 
573: perturbations that would arise from the back reaction.  However, these higher 
574: harmonics are suppressed by additional factors of $H \propto \Psi^2$: 
575: whereas the static and $\cos {2\omega t}$ components of $H$ are quadratic in 
576: $\Psi$, higher-order harmonics of frequency $2n\omega$ are proportional to 
577: $\Psi^{2n}$ for $n>1$.    
578: 
579: 
580: \subsection{\label{sec:sw-nonlin}Time-averaged fully nonlinear 
581: standing-wave solution}
582: 
583: To explore the behavior of the standing-wave spherical scalar field and the 
584: spherical metric near and inside the expected location of the horizon, we solve 
585: the fully nonlinear coupled Einstein-scalar-field equations including full 
586: back reaction in the wave equation.  To simplify our solution, we average the 
587: stress-energy tensor in time to remove oscillations of the scalar-field
588: energy, so that the metric is static.  This is justified by the perturbative 
589: analysis above, which demonstrates that metric components oscillating in time 
590: are smaller than static metric components and largely decouple from them.
591: 
592: \subsubsection{Formalism for nonlinear solution with back reaction}
593: We write the static spherically symmetric metric in the form
594: \begin{equation}\label{metricstat}
595: ds^2=-e^{\beta(r)+\alpha(r)}dt^2+e^{\beta(r)-\alpha(r)}dr^2+r^2d\Omega^2 ,
596: \end{equation}
597: and we compute the Einstein tensor from this metric in the standard way.  It is
598: diagonal and its angular components $G_{\hat{\theta}\hat{\theta}}$ and
599: $G_{\hat{\phi}\hat{\phi}}$ are not particularly interesting because
600: of spherical symmetry (the angular components of the Einstein equation will 
601: merely repeat the time and radial components by virtue of the 
602: contracted Bianchi identities).  The careted subscripts $ _{\hat{\mu}}$ denote
603: the orthonormal basis associated with the $(t, r, \theta, \phi)$ coordinate
604: system.  The relevant non-vanishing terms of the Einstein tensor in the
605: orthonormal basis are:
606: \begin{subequations}\label{G}
607: \begin{eqnarray}
608: G_{\hat{t}\hat{t}}&=&
609: 	e^{\alpha-\beta} (\beta'-\alpha')/r+(1-e^{\alpha-\beta})/r^2 ,
610: \\
611: G_{\hat{r}\hat{r}}&=&
612: 	e^{\alpha-\beta} (\beta'+\alpha')/r-(1-e^{\alpha-\beta})/r^2 ,
613: \end{eqnarray}
614: \end{subequations}
615: where $'$ denotes a derivative with respect to $r$, not $r^*$.
616: 
617: Substituting the Einstein tensor (\ref{G}) and the stress-energy tensor 
618: (\ref{T}) into the Einstein equations (\ref{Einstein}), we obtain: 
619: \begin{subequations}\label{Ein}
620: \begin{equation}\label{Ein1}
621: \alpha'=\frac{1}{r}(e^{\beta-\alpha}-1) \, ,
622: \end{equation}
623: \begin{equation}\label{Ein2}
624: \beta'= r e^{-2 \alpha} (\Phi_{,t})^2 + r (\Phi_{,r})^2 \, .
625: \end{equation}
626: \end{subequations}
627: 
628: We can now insert the standing-wave scalar-field ansatz (\ref{Phistat}) and 
629: time average the right hand side of Eq.~(\ref{Ein2}) over a complete period.
630: For numerical analysis it will be useful to switch to a logarithmic coordinate
631: that changes more gradually than $r$ in the vicinity of the Schwarzschild 
632: horizon.  The following generalization of the Regge-Wheeler tortoise coordinate
633: $r^*$ proves convenient:
634: \begin{subequations}\label{Dtot}
635: \begin{equation}\label{Dr}
636: \frac{d r}{d r^*} = e^{\alpha} \, .
637: \end{equation}
638: 
639: In terms of this coordinate, the wave equation (\ref{wave}) simplifies to
640: \begin{equation}\label{DPsi}
641: \frac{d^2 \Psi}{d {r^*}^2} = - \omega^2 \Psi + \frac{e^{\alpha}}{r} 
642: 	\frac{d\alpha}{d r^*} \Psi
643: \end{equation}
644: and the Einstein equations (\ref{Ein}) with time-averaged 
645: $(\Phi_{,t})^2$ and $(\Phi_{,r})^2$ become
646: \begin{equation}\label{Dalpha}
647: \frac{d \alpha}{d r^*}=\frac{e^{\beta}-e^{\alpha}}{r} \, ,
648: \end{equation}
649: \begin{equation}\label{Dbeta}
650: \frac{d \beta}{d r^*}=
651: \frac{e^{-\alpha}}{2r}\Biggl[\Psi^2 \omega^2 +\left(\frac{d\Psi}{dr^*}\right)^2
652: 	\Biggr]+
653: \frac{\Psi^2 e^{\alpha}}{2 r^3} -
654: \frac{\Psi}{r^2}\frac{d\Psi}{dr^*} \, .
655: \end{equation}
656: \end{subequations}
657: 
658: \subsubsection{Singular standing-wave spacetime}
659: We have solved the coupled equations (\ref{Dtot}) numerically to high
660: accuracy for values of the scalar-field amplitude and frequency in the range
661: relevant to the BBH problem [Eqs.~(\ref{arange})].  Our numerical solutions are
662: very well approximated by analytic formulae that rely on dividing space
663: $0 < r < \infty$ into three regions. Region I is  ``perturbed 
664: Schwarzschild'', i.e., the region where the perturbative solution is valid
665: ($r>2$, $H \alt 0.1$).  Region III describes the space very close to 
666: $r=0$ where the $1/r$ terms diverge.  Finally, the intermediate 
667: region II extends from the inner boundary of region I to the outer boundary 
668: of region III.
669: 
670: For sufficiently small amplitudes of the scalar field, the contributions
671: from the {\it back reaction} (by which we mean the impact of the deviation of
672: the spacetime from Schwarzschild on the solution to the wave equation) and from
673: nonlinearity remain small until very close to $r=2$, so that the metric can be 
674: well approximated by perturbations on top of the Schwarzschild metric.  In
675: other words, the perturbative solution developed in Sec.~\ref{sec:sw-pert} is
676: valid throughout region I.  Indeed, for scalar-field amplitudes and frequencies
677: in the range of interest, the metric perturbations $H_0^{stat}$ and 
678: $H_2^{stat}$ derived in the previous subsection match the values 
679: of $H_0$ and $H_2$ corresponding to the complete nonlinear solution with back 
680: reaction to within $3\%$ for $H \approx 0.01$.
681: 
682: We begin the analysis in region III, where $r \ll 1$, by assuming 
683: $e^{\beta-\alpha} \ll 1$ as $r \to 0$, which corresponds to $g_{rr} \to 0$ 
684: at $r=0$.  (This assumption, which can be deduced from the behavior of 
685: $d\beta / dr^*$ in the transition region, will be shown to be self-consistent; 
686: more importantly, it is supported by our numerical solutions.)  Then, from
687: Eq.~(\ref{Dalpha}), $\alpha' \equiv d\alpha/dr \to -1/r$, so $\alpha$ is given 
688: by
689: \begin{subequations}\label{tot3}
690: \begin{equation}\label{alpha3}
691: \alpha = - \ln r + \alpha_0 \, .
692: \end{equation}
693: Here $\alpha_0$ is a constant whose value depends on the amplitude 
694: and the frequency of the scalar waves; it can be roughly approximated by
695: \begin{equation}\label{alpha0}
696: \alpha_0 \sim  \ln{\left(A^2\omega^2\right)}\, ,
697: \end{equation}
698: where $A$ is the amplitude of the scalar field near $r=2$.
699: 
700: The wave equation (\ref{DPsi}) becomes
701: \begin{eqnarray}\label{PPsi3}
702: \Psi'' &=& - \Psi \omega^2 e^{-2\alpha} - \alpha'(\Psi'-\Psi/r) 
703: \\
704: \nonumber
705: &=&
706: -\Psi \omega^2 e^{-2\alpha_0} r^2 + 1/r (\Psi'-\Psi/r) \, .
707: \end{eqnarray}
708: Since we are interested in the region $r \to 0$, the last term dominates, so
709: that the approximate solution to Eq.~(\ref{PPsi3}) is
710: \begin{equation}\label{Psi3}
711: \Psi = n r + k r \ln r \, ,
712: \end{equation}
713: where $n, k$ are constants.
714: 
715: Substituting $\Psi$ and $\alpha$ into Eq.~(\ref{Dbeta}) and selecting 
716: non-vanishing terms with the highest order in $1/r$, we find that 
717: $\beta' \to k^2 /(2r)$, so
718: \begin{equation}\label{beta3}
719: \beta = \frac{k^2}{2} \ln {r} + \beta_0 \, ,
720: \end{equation}
721: where $\beta_0$ is a constant.
722: Thus, we see that our assumption, $e^{\beta-\alpha} \ll 1$ as $r \to 0$, is 
723: self-consistent:
724: \begin{equation}\label{ba3}
725: \beta-\alpha = \left(\frac{k^2}{2}+1\right) \ln r +\beta_0 -\alpha_0  
726: 	\to -\infty \, \text{as} \, r \to 0 ,
727: \end{equation}
728: since the coefficient of $\ln r$ is always positive.
729: 
730: Our numerical solution in region III agrees well with the asymptotic 
731: analytical solution (\ref{tot3}).  For instance, the value of $k$ obtained 
732: from matching $\Psi$ to the form of Eq.~(\ref{Psi3}) agrees with the value 
733: of $k$ obtained from matching $\beta$ to Eq.~(\ref{beta3}) to one part in ten 
734: thousand.  Of particular interest are the metric components and the
735: Ricci scalar, whose asymptotics for $r \to 0$ are:
736: \begin{equation}\label{gtt3}
737: g_{tt}=-e^{\beta+\alpha}=-e^{\beta_0+\alpha_0} r^{k^2/2 - 1} \, ,
738: \end{equation}
739: \begin{equation}\label{grr3}
740: g_{rr}=e^{\beta-\alpha}=e^{\beta_0-\alpha_0} r^{k^2/2 + 1} \, ,
741: \end{equation}
742: and
743: \begin{equation}\label{Rscalar3}
744: R=R_{\gamma}^{\gamma}=2\Phi_{,\gamma}\Phi^{,\gamma}=k^2 
745: 	e^{\alpha_0-\beta_0} r^{-3-k^2/2} \, .
746: \end{equation}
747: \end{subequations}
748: The exponent of $r$ in Eq.~(\ref{Rscalar3}) is always negative, so the
749: Ricci curvature scalar tends to infinity as $r \to 0$, i.e., the radius of 
750: curvature vanishes at the singularity at $r=0$, as expected. 
751: The exponent of $r$ in Eq.~(\ref{grr3}) is always positive, so
752: $g_{rr}$ tends to zero as $r \to 0$ according to a power law.  However,
753: the sign of the exponent of $r$ in Eq.~(\ref{gtt3}) depends on the value
754: of $k$, which in turn is a complicated function of the scalar-field 
755: frequency and amplitude.  For some scalar field parameter values in 
756: the range relevant to the BBH problem [Sec.~\ref{sec:sfeq}] $k^2/2 > 1$
757: and $g_{tt}$ vanishes at the singularity; for others, $g_{tt}$ 
758: is infinite at $r=0$.
759: 
760: 
761: The nature of region II, which represents the transition from the 
762: Schwarzschild-like region I to the singularity of region III, depends strongly 
763: on the values of $\Psi_0$ and $\omega$.  In Schwarzschild, 
764: $\alpha=\ln{(1-2/r)}$ tends to $-\infty$ as $r \to 2$, and this is the 
765: behavior of $\alpha$ in the nearly Schwarzschild region I; meanwhile, in 
766: region II, as in region III, $\alpha$ is well approximated by
767: \begin{subequations}\label{tot2}
768: \begin{equation}\label{alpha2}
769: \alpha = - \ln r + \alpha_0 \, .
770: \end{equation}
771: The outer boundary of region II is located at the transition between these 
772: two behaviors of $\alpha$, i.e., at the minimum of $\alpha$.  
773: 
774: Substituting the approximation (\ref{alpha2}) for $\alpha$ into the wave 
775: equation (\ref{DPsi}), we obtain:
776: \begin{equation}\label{DPsi2}
777: \frac{d^2\Psi}{d {r^*}^2}=\Psi \bigl(-\omega^2 - \frac{e^{2\alpha_0}}{r^4}\bigr)
778: \, .
779: \end{equation}
780: Thus, the condition for the scalar field to exhibit spatial oscillations at 
781: an approximately constant amplitude is 
782: $e^{2\alpha_0} / r^4 \ll \omega^2$.  The location where this condition
783: begins to be violated forms the inner boundary of region II.  Thus, region II
784: can be said to be defined by the variation of $\alpha$ according to
785: Eq.~(\ref{alpha2}) as in region III, and by rapid spatial oscillations of 
786: the scalar field $d\Psi / dr^* = \omega$ as in region I.
787: 
788: Since $\alpha_0$ will be more negative for smaller amplitudes of the scalar 
789: field, we see that region II is going to be significant for small $\Psi_0$,
790: including those in the relevant range of the BBH problem.  For larger values 
791: of $\Psi_0$, the metric and scalar field will proceed directly from region 
792: I to region III.
793: 
794: When region II does exist, the amplitude and phase of the scalar field 
795: [solution of Eq.~(\ref{DPsi2})]
796: \begin{equation}\label{Psi2}
797: \Psi(r) = A(r) \cos {\phi(r)}
798: \end{equation}
799: will be given by
800: \begin{equation}\label{A2}
801: A = A_0 (1 - \frac{e^{2\alpha_0}}{4 r^4 \omega^2} + ...) \, ,
802: \end{equation}
803: \begin{equation}\label{phi2}
804: \dot{\phi}=\omega (1 + \frac{e^{2\alpha_0}}{2 r^4 \omega^2} + ...) \, ,
805: \end{equation}
806: to first order in $e^{2\alpha_0}/(r^4 \omega^2)$.
807: 
808: Substituting expressions (\ref{tot2}) for $\alpha$ and $\Psi$ into the 
809: differential equation for $\beta$, Eq.~(\ref{Dbeta}), we find that the dominant 
810: term is the first one,
811: $d\beta / dr^* \to (1/2) e^{-\alpha_0} A^2 \omega^2$, so in region II
812: $\beta$ is approximately
813: \begin{eqnarray}
814: \nonumber
815: \beta &=& \frac{1}{2} e^{-\alpha_0} A^2 \omega^2 r^* + \text{const} 
816: \\
817: \label{beta2}
818: 	&=& \frac{1}{4} e^{-2\alpha_0} r^2 A^2 \omega^2 +\text{const} \, ,
819: \end{eqnarray}
820: where the last equality comes from the integral of equation (\ref{Dr}),
821: $r^* = e^{-\alpha_0} r^2 / 2 + \text{const}$.
822: \end{subequations}
823: 
824: Embedding diagrams and redshifts may provide the best pictorial insight into 
825: our full time-averaged standing-wave scalar-field solution, including all of
826: regions I, II and III.  
827: 
828: Figure \ref{emb-fig} shows an embedding diagram for the standing-wave spacetime:
829: \begin{equation}\label{emb}
830: \frac{dz}{dr} = \sqrt{|g_{rr}-1|}
831: \end{equation}
832: The 2-surface obtained by rotating around the vertical axis $r=0$
833: has the same 2-geometry as the surface $(t=\text{const},\, \theta=\pi/2)$ in
834: the standing-wave spacetime.  At radii $r>2$ the embedding is very nearly the
835: same as for a Schwarzschild black hole (cf.~Fig.~$31.5$ of \cite{MTW}).   
836: For $r<2$, the radial distance changes far more slowly than the areal radius 
837: ($0<g_{rr} \ll 1$), so the embedding is performed in Minkowski space rather
838: than Euclidean space: the metric is $ds^2=-dz^2+dr^2+r^2d\phi^2$ rather than 
839: $ds^2=+dz^2+dr^2+r^2d\phi^2$.  The embedded surface asymptotes to the light cone
840: as $r\to0$. 
841: 
842: Figure \ref{red-fig} depicts the redshift of light emitted at one radius
843: and received at another, greater one, as a function of the emitting radius:
844: \begin{equation}\label{red}
845: z=\sqrt{\frac{g_{tt}^{rec}}{g_{tt}^{em}}}-1
846: \end{equation}
847: Figure \ref{reda-fig} shows that, while the redshift becomes very large as 
848: $r \rightarrow 2$, it never becomes infinite there as it would for a 
849: Schwarzschild black hole.  As expected, the horizon is destroyed by 
850: the standing-wave scalar field, so an observer at infinity can receive 
851: signals from any source at $r>0$, albeit with a very large redshift for 
852: sources close to or inside the location ($r=2$) of the Schwarzschild 
853: horizon.  A blown-up view of the region $r \ll 1$ [Figure \ref{redb-fig}] 
854: shows that the signal emitted near the singularity may
855: be infinitely redshifted or blueshifted depending on the asymptotics of
856: the scalar field as $r \rightarrow 0$ according to
857: \begin{equation}\label{red3}
858: z=\sqrt{g_{tt}^{rec}} e^{(\alpha_0-\beta_0)/2} r^{-k^2/4+1/2} - 1 \, . 
859: \end{equation}
860: 
861: 
862: \begin{figure}[ht]
863: \includegraphics[keepaspectratio=true,width=8.5cm,angle=0]{emb.eps}
864: \caption{Embedding diagram for the spacetime with time-averaged standing-wave 
865: scalar field of angular frequency 
866: $\omega=0.19$ and amplitude $\Psi_0=0.015$ at large radii 
867: [corresponding to the binary black hole separation $a \approx 6M$;
868: Eq.~(\ref{a6})].  The solid line represents embedding in Euclidean space;
869: the dashed line, embedding in Minkowski space.  Regions I, II and III 
870: are labeled on plot.}
871: \label{emb-fig}
872: \end{figure}
873: 
874: \begin{figure}[ht]
875: \subfigure{
876: \includegraphics[keepaspectratio=true,width=8.5cm,angle=0]{reda.eps}
877: \label{reda-fig}
878: }
879: \subfigure{
880: \includegraphics[keepaspectratio=true,width=8.5cm,angle=0]{redb.eps}
881: \label{redb-fig}
882: }
883: \caption{(a) Redshift $z=\delta \lambda / \lambda$ of light emitted from
884: radius $r$ and received by an observer at $r=10$.  (b) Redshift for an observer 
885: at  $r=0.0001$.  A distant observer would see light emitted from
886: $r=0.0001$ redshifted by $\ln(z+1) \approx 10^5$.
887: These curves are drawn for the spacetime with time-averaged standing-wave 
888: scalar field that has angular frequency 
889: $\omega=0.19$ and amplitude $\Psi_0=0.015$ at large radii 
890: [corresponding to the binary black hole separation $a \approx 6M$;
891: Eq.~(\ref{a6})].}
892: \label{red-fig}
893: \end{figure}
894: 
895: \subsubsection{Comparison of standing-wave and Schwarzschild spacetimes}
896: It is important to understand how the complete standing-wave spacetime 
897: with back reaction (we shall call this spacetime $S$) compares with
898: the Schwarzschild spacetime (which we shall call spacetime $D$).  We might
899: first try to compare the metric components in the two spacetimes as functions
900: of the radial coordinate $r$.  Indeed, the metric components 
901: $g_{\theta\theta}=r^2$ and $g_{\phi\phi}=r^2\sin^2{\theta}$ are precisely 
902: equal in the two spacetimes when evaluated at the same location in 
903: $(t, r, \theta, \phi)$ coordinates.
904: Furthermore, outside the effective-potential region, the perturbation 
905: due to the scalar field is so small that the fractional difference 
906: $\delta g_{\alpha\beta}/g_{\alpha\beta} \equiv 
907: (g^S_{\alpha\beta}-g^D_{\alpha\beta})/g^D_{\alpha\beta}$ in metric 
908: components $g_{tt}$ and $g_{rr}$ does not exceed $0.01\%$
909: for scalar-field parameters in the range of interest.
910: However, the metric components $g_{tt}$ and $g_{rr}$ in $S$ and $D$ can differ 
911: by orders of magnitude near $r=2$, inside the effective potential peak.  
912: 
913: The apparent mismatch between the metric components of the two spacetimes near
914: $r=2$ turns out to be a consequence of a poor choice of the radial 
915: coordinate $r$ for comparison.  A much better choice is $r^*$:  
916: when the coordinates $(t, r^*, \theta, \phi)$ are used for mapping between
917: the two spacetimes $S$ and $D$, the metric components agree extremely
918: well near $r=2$.
919: 
920: The fractional differences $\delta g / g$ between 
921: the $g_{tt}$ and $g_{\theta\theta}$ components in $S$ and $D$ are 
922: plotted in Fig.~\ref{spacetime-fig} for scalar field parameters corresponding 
923: to binary black hole separations at the boundaries of the range of interest.  
924: Using $r^*$ rather than $r$ as the coordinate for comparison means that 
925: the $g_{\theta\theta}$ components no longer match perfectly; however, the
926: fractional difference introduced remains small as $r \rightarrow 2$ and 
927: does not exceed $0.6\%$ in the range of interest.  
928: The fractional differences in $g_{\phi\phi}$ are identical to those in 
929: $g_{\theta\theta}$ and are not plotted separately.  
930: The Regge-Wheeler tortoise coordinate $r^*$ 
931: [Eq.~(\ref{rRW})] and its generalization [Eq.~(\ref{Dr})] were defined
932: so that $g_{r^*r^*} \equiv -g_{tt}$ in both spacetimes $S$ and $D$; therefore,
933: the fractional differences in the values of $g_{r^*r^*}$ in $S$ and $D$ are 
934: the same as the fractional differences in $g_{tt}$.
935: 
936: As Fig.~\ref{spacetime-fig} shows, the fractional differences in the metrics 
937: are $\lesssim 0.02$ down to values of $r^* \sim -1000$, a location so deep 
938: inside the peak of the effective potential that it contains  at least $20$ 
939: near-horizon oscillations of the scalar field for frequencies and 
940: amplitudes in the BBH separation range of interest.  
941: Perhaps a more impressive way to state this is that 
942: in the $(t, r^*, \theta, \phi)$  coordinate system, metric 
943: components of $g^S$ and  $g^D$ match to an accuracy of $2\%$ 
944: for all relevant scalar-field parameters down to the Schwarzschild radius 
945: $r^D - 2 < 10^{-100}$.  
946: 
947: The fractional differences between the coefficients of the metrics $g^S$ and 
948: $g^D$ continue to grow approximately linearly in $r^*$ deep inside the
949: effective potential and reach $10\%$ at the Schwarzschild radius
950: $r^D - 2 \sim 10^{-3000}$, or approximately 500 scalar-field 
951: oscillations inside the effective-potential peak for scalar field parameters
952: corresponding to BBH separation $a \approx 6 M$.
953: 
954: \begin{figure}[ht]
955: \subfigure{
956: \includegraphics[keepaspectratio=true,width=8.5cm,angle=0]{SvsD6.eps}
957: \label{spacetime6-fig}
958: }
959: \subfigure{
960: \includegraphics[keepaspectratio=true,width=8.5cm,angle=0]{SvsD15.eps}
961: \label{spacetime15-fig}
962: }
963: \caption{(a)Fractional differences of the metric components 
964: $g_{tt}=-g_{r^*r^*}$ (solid curve) and $g_{\theta\theta}$ (dashed curve) 
965: between Schwarzschild spacetime $D$ and standing-wave scalar field spacetime 
966: $S$ with scalar-wave amplitude and frequency chosen to model BBH separation 
967: $a \approx 6M$ [Eq.~(\ref{a6})].
968: (b)Same quantities plotted for scalar field parameters chosen to model
969: BBH separation $a \approx 15M$ [Eq.~(\ref{a15})].}
970: \label{spacetime-fig}
971: \end{figure}
972: 
973: 
974: 
975: \section{\label{sec:down}Downgoing scalar field}
976: 
977: Having discussed, in Sec.~\ref{sec:sw}, the standing-wave scalar-field 
978: spacetime that modeled the stationary BBH approximation, we now turn to a
979: scalar-field spacetime that  serves as a model for the physical BBH spacetime
980: with downgoing gravitational waves at the black-hole horizons:
981: a nearly Schwarzschild spacetime with spherically symmetric scalar waves that 
982: are purely downgoing at $r=2$.  
983: 
984: For a perturbative analysis of downgoing scalar waves in Schwarzschild, the 
985: ingoing, null Eddington-Finkelstein time coordinate $v=t+r^*$ is more 
986: appropriate than the standard Schwarzschild time coordinate.  Let us suppose 
987: that by the time $v=0$ a total mass-energy $M_0 = m(v=0)$ is located within 
988: the horizon $r=2$.  We are not particularly interested in how this mass 
989: accumulated there or how the scalar field behaved in the past; we are only 
990: interested in the times immediately following $v=0$, and we let the scalar 
991: waves be purely downgoing and monochromatic at the horizon for $v>0$.  Then for 
992: $v>0$ radiation is falling into the black hole at a nearly constant rate, 
993: corresponding to the energy density in the scalar field 
994: $dm/dv \approx \Psi_0^2 \omega^2 / 2$, with some small oscillations 
995: on top of the linear increase in mass.  This is very similar to the Vaidya
996: solution and, indeed, the Vaidya metric will be seen to describe the 
997: spacetime of the downgoing scalar-field solution:
998: \begin{equation}\label{vaidya}
999: ds^2 = -\left[1-\frac{2m(v)}{r}\right] dv^2 + 2dv dr + r^2 d\Omega^2 \, .
1000: \end{equation}
1001: 
1002: Near the horizon, $\Phi=(1/r)\cos{\omega v}$ is a purely downgoing 
1003: solution to the wave equation (\ref{wave}).  The only non-zero term of the 
1004: Ricci tensor in Vaidya coordinates is $R_{vv}=(2/r^2) m'(v)$, where
1005: $'$ denotes the derivative with respect to $v$.  The Einstein 
1006: equations (\ref{Ricci}) at $r=2$ say:
1007: \begin{equation}\label{Riccidown}
1008: R_{vv}=\frac{2 m'(v)}{4} = 2 \Phi_{,v} \Phi_{,v} = 
1009: 	\frac {2 \Psi_0^2 \omega^2 \sin^2{\omega v}}{4} \, .
1010: \end{equation}
1011: 
1012: Equation (\ref{Riccidown}) is trivially integrated to obtain:
1013: \begin{equation}\label{mdown}
1014: m(v) = M_0 + \frac{\Psi_0^2 \omega^2}{2} v - 
1015: 	\frac{\Psi_0^2 \omega \sin{2\omega v}}{4} \, .
1016: \end{equation}
1017: The black-hole mass grows linearly in $v$ at the rate $\Psi_0^2 \omega^2/2$
1018: with a tiny superimposed oscillatory component.  The black hole retains a
1019: smooth, non-singular future horizon.
1020: 
1021: The scalar field is purely downgoing at the horizon and approximately downgoing 
1022: everywhere inside the Schwarzschild effective-potential peak.  Outside the 
1023: effective-potential peak there is both a downgoing scalar field and an upgoing 
1024: one, reflected off the potential.  Since for small $v$ the metric is nearly 
1025: Schwarzschild [the constant Schwarzschild mass $M$ is replaced by the $m(v)$ 
1026: of Eq.~(\ref{mdown})], the scalar field everywhere is given to a high accuracy 
1027: by a solution to the wave equation in the Schwarzschild background subject to 
1028: the purely downgoing boundary condition at the horizon.  (Of course, very far 
1029: from the horizon the energy contained in the intervening scalar field will act 
1030: as an additional mass, but we are not interested in this region for our model 
1031: problem.)
1032: 
1033: 
1034: \section{\label{sec:reconstruct}Reconstruction of downgoing scalar field 
1035: from standing-wave scalar field}
1036: 
1037: We turn now to our scalar-wave version of adding a radiation-reaction
1038: field to a standing-wave spacetime to obtain a physical spacetime with
1039: downgoing waves at horizons and outgoing waves at infinity.  For this
1040: procedure there is a substantial difference between the BBH problem and our
1041: model problem.
1042: 
1043: In the true BBH problem, the periodic standing wave (SW) solution is sourced
1044: by the black holes and corresponds to the $\frac{1}{2} \text{Retarded} + 
1045: \frac{1}{2} \text{Advanced}$ solution of the Green's function problem.  In 
1046: this case we add the non-sourced $\frac{1}{2} \text{Retarded} - 
1047: \frac{1}{2} \text{Advanced}$ radiation reaction (RR) solution of the linearized
1048: Einstein equations in the SW spacetime to get an approximation to the 
1049: physical retarded solution \cite{private}.  At infinity, where the SW 
1050: field is $\frac{1}{2} \text{Outgoing} + \frac{1}{2} \text{Ingoing}$, the
1051: boundary condition for the RR field should be set to
1052: $\frac{1}{2} \text{Outgoing} - \frac{1}{2} \text{Ingoing}$, so
1053: that their sum contains only physical outgoing waves, and similarly at the 
1054: horizons the RR field will be 
1055: $\frac{1}{2} \text{Downgoing} - \frac{1}{2} \text{Upgoing}$.  Adding this 
1056: RR field to the $\frac{1}{2} \text{Downgoing} + 
1057: \frac{1}{2} \text{Upgoing}$ standing waves would yield gravitational waves
1058: that are downgoing at the expected horizon locations, conforming
1059: to the expected behavior in physical black-hole spacetimes.  (We do not
1060: expect the stress-energy tensor of the sum of SW and RR waves to precisely 
1061: match the Einstein tensor of the SW spacetime because, of course, gravitational 
1062: theory is not linear; however, it is likely that "effective linearity" holds 
1063: in the sense defined by Price \cite{Price} for the non-tidally-locked case as 
1064: well as for the tidally-locked case.  In a future paper we intend to explore
1065: this issue with a model that more closely resembles the BBH problem.)
1066: 
1067: The scalar-field model we are currently analyzing is not sourced:
1068: the wave equation (\ref{wave}) we used to compute the SW solution 
1069: is homogeneous.  There is then no perturbative homogeneous
1070: solution that is simultaneously $\frac{1}{2} \text{Outgoing} - 
1071: \frac{1}{2} \text{Ingoing}$ at infinity and $\frac{1}{2} \text{Downgoing} 
1072: - \frac{1}{2} \text{Upgoing}$ at the expected horizon location.  
1073: Since at the outer boundary the problem is obviously linear for 
1074: sufficiently weak scalar fields, it is easy to reconstruct the outgoing 
1075: solution from the SW solution there: we simply double the outgoing 
1076: component of the SW solution.  The interesting case lies in the extraction 
1077: of a downgoing solution near $r \approx 2$.  We attempt to reconstruct the
1078: downgoing scalar field from the SW scalar field near the expected horizon
1079: by adding to the SW field a perturbative RR field that is
1080: $\frac{1}{2} \text{Downgoing} - \frac{1}{2} \text{Upgoing}$ at 
1081: $r \approx 2$.
1082: 
1083: As in Sec.~\ref{sec:sw-nonlin}, let $S$ denote the spacetime of the complete 
1084: standing-wave solution with back reaction.  As discussed in the previous
1085: section, the spacetime of the downgoing scalar field is approximated to
1086: sufficient accuracy for our purposes by the Schwarzschild spacetime $D$.
1087: 
1088: The complete SW scalar field is a solution to the wave equation in 
1089: spacetime $S$ (in our simplified treatment of the problem, spacetime $S$ 
1090: actually corresponds to the time-averaged solution, i.e., one in which we 
1091: ignore the oscillatory components of the metric). The RR field is a solution 
1092: to the same wave equation in $S$ in our model.  The ``reconstructed'' downgoing 
1093: field is, therefore, the downgoing solution to the wave equation in $S$.  
1094: We want to compare this to the ``true'' downgoing field, which is the downgoing 
1095: solution to the wave equation in $D$, i.e., in Schwarzschild. 
1096:  
1097: \begin{figure}[tbh!]
1098: \subfigure{
1099: \includegraphics[keepaspectratio=true,width=8.5cm,angle=0]{recon6.eps}
1100: \label{recona-fig}
1101: }
1102: \subfigure{
1103: \includegraphics[keepaspectratio=true,width=8.5cm,angle=0]{recon15.eps}
1104: \label{reconb-fig}
1105: }
1106: \caption{(a) The fractional difference in the amplitudes of the reconstructed 
1107: scalar field and downgoing scalar field 
1108: $\delta d / d \equiv (d^{SW+RR}-d^{down})/d^{down}$
1109: (solid curve) and the phase difference between the 
1110: two fields $\delta \phi_d = \phi_d^{SW+RR}-\phi_d^{down}$ (dashed curve),
1111: plotted vs. $r^*$.  Scalar-wave amplitude and frequency chosen to model BBH
1112: separation $a \approx 6M$.
1113: (b) Same quantities plotted for scalar-field parameters chosen to model 
1114: BBH separation $a \approx 15M$.}
1115: \label{recon-fig}
1116: \end{figure} 
1117:  
1118: In the region between the expected horizon location $r=2$ and the inner edge 
1119: of the peak of the effective potential, the wave equation (\ref{wave}) is 
1120: dominated by 
1121: \begin{equation}\label{appwave}
1122: \frac{d^2\Psi}{d{r^*}^2} \approx -\omega^2 \Psi 
1123: \end{equation}
1124: in both spacetimes $S$ and $D$.  Hence, the solution to the wave equation
1125: will be oscillatory in $r^*$ with frequency $\omega$, which makes
1126: sense on physical grounds, since ingoing light cones are 
1127: $t+r^*=\text{constant}$ in both $S$ and $D$.  Moreover, as discussed 
1128: in  Sec.~\ref{sec:sw-nonlin}, the metrics of the two spacetimes
1129: are nearly the same in the $r^*$ coordinate, i.e., $g^S(r^*) \approx g^D(r^*)$. 
1130: This suggests that to get the scalar wave phasing to agree, we need
1131: to map between the two spacetimes using the $r^*$ radial coordinate.
1132: 
1133: We set the boundary conditions for both the RR field in $S$ and the downgoing
1134: field in $D$ at a negative value of $r^*$ chosen so that the fields are
1135: at least a few wavelengths inside the effective potential, and so that 
1136: $r^\text{S}(r^*)$ is very close to $r^\text{S}=2$ (it might actually be 
1137: slightly inside $r=2$).  The $\text{SW}+\text{RR}$ and downgoing scalar fields 
1138: will match by construction at the point where the initial conditions are set.  
1139: We will integrate both solutions toward larger $r^*$ and compare the quality of
1140: the match between the two fields.
1141: 
1142: For the purposes of comparing the scalar fields in the two spacetimes,
1143: we separate the complex scalar field $\Psi(r^*)$ [the spatial factor of 
1144: the complete field $\Phi(r,t)=\Re[\Psi(r^*) e^{-i\omega t}]/r$, 
1145: cf.~Eq.~(\ref{Phi})] into upgoing and downgoing components.  We define
1146: the amplitudes and phases of the upgoing and downgoing fields as follows (see
1147: below for motivation):
1148: \begin{subequations}\label{PsiDU}
1149: \begin{eqnarray}
1150: u&\equiv&\frac{1}{2\omega} \left| \frac{d\Psi}{dr^*}+i\omega\Psi\ \right| ;
1151: \\
1152: d&\equiv&\frac{1}{2\omega} \left| \frac{d\Psi}{dr^*}-i\omega\Psi\ \right| ;
1153: \\
1154: e^{i\phi_u}&\equiv&\frac{1}{2i\omega u} 
1155: 	\left( \frac{d\Psi}{dr^*}+i\omega\Psi\ \right) ;
1156: \\
1157: e^{i\phi_d}&\equiv&\frac{1}{-2i\omega d} 
1158: 	\left( \frac{d\Psi}{dr^*}-i\omega\Psi\ \right) \, .
1159: \end{eqnarray}
1160: \end{subequations}
1161: 
1162: To motivate these definitions we consider the
1163: geometric optics limit, where the wave phase evolves much 
1164: faster than the amplitude.  In this limit,
1165: the downgoing component of the scalar field 
1166: $\Psi_d \propto e^{-i\omega r^*}$ separates unambiguously
1167: from the upgoing component $\Psi_u \propto e^{i\omega r^*}$:
1168: \begin{subequations}\label{PsiDUmot}
1169: \begin{equation}\label{Psicomp}
1170: \Psi(r^*) =  u e^{i \phi_u} + d e^{i \phi_d} \, ,
1171: \end{equation}
1172: where we use the standard approximations
1173: \begin{eqnarray}\label{phiu}
1174: \frac{d \phi_u}{dr^*} &\cong& \omega \gg \left| \frac{d u}{dr^*}\right| ,
1175: \\
1176: \label{phid}
1177: - \frac{d \phi_d}{dr^*} &\cong& \omega \gg \left|\frac{d d}{dr^*}\right| \, .
1178: \end{eqnarray}
1179: \end{subequations}
1180: Inverting Eq.~(\ref{Psicomp}) with these approximations yields the definitions
1181: (\ref{PsiDU}).
1182: Although the geometric optics approximations break down in the region of 
1183: the effective potential, and the separation of the scalar waves into 
1184: upgoing and downgoing components becomes ambiguous there because the wave
1185: speed is ill-determined outside the short-wavelength limit, expressions
1186: (\ref{PsiDU}) are adequate for comparing scalar fields in our region
1187: of interest.  
1188: 
1189: In Fig.~\ref{recon-fig} we show the fractional difference 
1190: $\delta d / d \equiv (d^{SW+RR}-d^{down})/d^{down}$ in the amplitude of 
1191: the downgoing components of the reconstructed $\text{SW}+\text{RR}$ waves 
1192: and the downgoing waves along with the phase difference 
1193: $\delta \phi_d = \phi_d^{SW+RR}-\phi_d^{down}$.
1194: The two plots represent the endpoints of the range of relevant BBH 
1195: separations: $a\approx 6M$ in Fig.~\ref{recona-fig} and 
1196: $a\approx 15M$ in Fig.~\ref{reconb-fig}.  Only the downgoing 
1197: amplitude $d$ and downgoing phase $\phi_d$ are plotted.  The upgoing field 
1198: components are zero to numerical precision inside the effective potential 
1199: and the differences between the reflected upgoing components of the 
1200: ``reconstructed'' and ``true'' downgoing fields outside the effective-potential 
1201: peak are similar to the differences between the downgoing field components 
1202: there, $\delta u/u \sim \delta_d/d$ and $\delta \phi_u \sim \delta \phi_d$.
1203: 
1204: The amplitudes and phases of the ``true'' downgoing field and the 
1205: ``reconstructed'' downgoing field match to within one part in ten million
1206: from the location where the initial conditions are set 
1207: (several scalar-field oscillations inside the effective potential) 
1208: to the inner edge of the effective-potential peak for all BBH separations in 
1209: the range of interest.  Near the effective-potential peak the 
1210: fractional difference in the amplitudes does not exceed $0.03\%$ and the phase 
1211: difference is less than $0.002$.  Outside the effective potential, the 
1212: fractional difference in the amplitudes is $5$ parts per million and the phase 
1213: difference is less than $0.00002$ for the smallest BBH separations in the range
1214: of interest.  
1215: 
1216: We also compared the ``reconstructed'' and ``true'' downgoing fields 
1217: very deep inside the effective potential when the field-matching initial
1218: conditions are set about 10 scalar-field oscillations inside the 
1219: effective-potential peak.  In this case, the amplitudes of the two fields are 
1220: equal to within numerical precision and the phase difference does not exceed 
1221: $3 \times 10^{-7}$ down to $500$ scalar-field oscillations inside the 
1222: effective-potential peak.  The fields begin to disagree significantly only 
1223: once the naked singularity is approached in the spacetime $S$, at
1224: $r^S(r^*) \lesssim 0.2$ 
1225: \footnote{It might seem odd that the fields continue to match far deeper 
1226: (at far more negative $r^*$) than the metrics, which begin to disagree by 
1227: $10\%$ at $500$ scalar-field oscillations inside the
1228: effective-potential peak.  The reason is that in the wave equation, the metric 
1229: enters only into the effective-potential piece [see Eq.~(\ref{DPsi})], 
1230: which is so tiny throughout the region $0.2 \lesssim r^S \lesssim 2$ that even 
1231: significant deviations of the standing-wave spacetime metric $g^S$ from the
1232: Schwarzschild metric do not affect the behavior of the scalar field.}.
1233: 
1234: 
1235: 
1236: 
1237: \begin{acknowledgments}
1238: I am very grateful to my advisor, Kip S. Thorne, for posing this research 
1239: project and for his extremely patient guidance.  Jonathan Gair, Lee Lindblom, 
1240: Robert Owen, Frans Pretorius and Mark A. Scheel of the TAPIR group at Caltech 
1241: made many helpful suggestions, as did Ronald J. Adler and Alexander S. 
1242: Silbergleit of the GP-B theory seminar at Stanford.  I was partially supported 
1243: by NSF Grant PHY-0099568 and NASA Grant NAG5-12834.
1244: \end{acknowledgments}
1245: 
1246: 
1247: \begin{thebibliography}{99}
1248: \bibitem{GravRad}
1249: K. S. Thorne, in 300 Years of Gravitation, edited by S. Hawking 
1250: and W. Israel (Cambridge University Press, 1987)
1251: \bibitem{Brady}
1252: P. R. Brady, J. D. E. Creighton, and K. S. Thorne, Phys. Rev. D
1253: \textbf{58}, 61501 (1998)
1254: \bibitem{Detweiler}
1255: S. Detweiler, Phys. Rev. D \textbf{50}, 4929 (1994)
1256: \bibitem{Price}
1257: R. Price, Class Quant. Grav. \textbf{21}, S281-S293 (2004);
1258: Z. Andrade \textit{et al.}, Phys. Rev. D \textbf{70}, 064001 (2004)
1259: \bibitem{private}
1260: K. S. Thorne, unpublished notes, 2002
1261: \bibitem{Wald}
1262: H. Friedrich, I. R\'acz, and R. M. Wald, Commun. Math. Phys.
1263: \textbf{204}, 691 (1999)
1264: \bibitem{MTW}
1265: C. W. Misner, K. S. Thorne, and J. A. Wheeler, \textit{Gravitation} 
1266: (Freeman, San Francisco, 1973)
1267: \bibitem{NF}
1268: V. P. Frolov and I. D. Novikov, \textit{Black Hole Physics}, 
1269: Fundamental Theories of Physics (Kluwer, Dordrecht, 1998)
1270: %\bibitem{GW}
1271: %K. Thorne, GW book, unpublished
1272: \bibitem{Poisson}
1273: E. Poisson and M. Sasaki, Phys. Rev. D \textbf{51}, 5753 (1995)
1274: \bibitem{Hua}
1275: H. Fang and G. Lovelace, in preparation
1276: \bibitem{RW}
1277: T. Regge and J. A. Wheeler, Phys. Rev. \textbf{108}, 1063 (1957)
1278: %\bibitem{Alvi}
1279: %K. Alvi, Phys. Rev. D \textbf{64}, 104020 (2001)
1280: \end{thebibliography}
1281: 
1282: \end{document}
1283: