1: %\documentclass[twocolumn,aps,prd,eqsecnum,nofootinbib]{revtex4}
2: %\documentclass[aps,prd,eqsecnum,nofootinbib]{revtex4}
3: \documentclass[10pt]{iopart}
4: \usepackage{epsfig}
5: %\usepackage[perpage,symbol*]{footmisc}
6: %\usepackage{amssymb}
7: %\usepackage{amsmath}
8: %\usepackage{amsfonts}
9: %\usepackage{bm}
10: \usepackage{iopams}
11: %\usepackage{psfrag}
12: \def\nn{\nonumber}
13: \def\no{\noindent}
14: \def\ve{\varepsilon}
15: \def\be{\begin{equation}}
16: \def\ee{\end{equation}}
17: \def\bfzeta{\mbox{\boldmath $\zeta$}}
18: \def\bfnabla{\mbox{\boldmath $\nabla$}}
19: \def\apj{Ap. J.}
20: \def\Phix{\phi}
21:
22: \eqnobysec
23:
24: \def\alt{\mathrel{\hbox{\rlap{\hbox{\lower4pt\hbox{$\sim$}}}\hbox{$<$}}}}
25: \def\agt{\mathrel{\hbox{\rlap{\hbox{\lower4pt\hbox{$\sim$}}}\hbox{$>$}}}}
26: %\def\Box{Box}
27:
28: %\DeclareMathSymbol{\R}{\mathbin}{AMSb}{"52}
29:
30: %\psfrag{scri+}{\LARGE{${\cal J}^{+}$}\normalsize}
31: %\psfrag{scri-}{\LARGE{${\cal J}^{-}$}\normalsize}
32:
33: \begin{document}
34:
35:
36: \title[Computing inspirals in Kerr in the adiabatic regime]{Computing inspirals in Kerr in the adiabatic regime. I. The scalar case}
37:
38: \author{Steve Drasco\dag, \'{E}anna \'{E}.\ Flanagan\dag\ddag, Scott A. Hughes\S
39: $\|$}
40:
41: \address{\dag\ Center for Radiophysics and Space Research, Cornell
42: University, Ithaca, NY 14853}
43: \address{\ddag\ Laboratory for Elementary Particle Physics, Cornell
44: University, Ithaca, NY 14853}
45: \address{\S\ Department of Physics, MIT, 77 Massachusetts Ave.,
46: Cambridge, MA 02139}
47: \address{$\|$\ MIT Kavli Institute for Astrophysics
48: and Space Research, MIT, 77 Massachusetts
49: Ave., Cambridge, MA 02139}
50:
51: \date{\today}
52: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
53: \begin{abstract}
54:
55: A key source for LISA will be the inspiral of compact objects
56: into massive black holes.
57: Recently Mino has shown that in the adiabatic limit, gravitational
58: waveforms for these sources
59: can be computed by using for the radiation reaction force the gradient of one
60: half the difference between the retarded and advanced metric
61: perturbations.
62: Using post-Newtonian expansions, we argue that the
63: resulting waveforms should be sufficiently accurate for signal detection with LISA.
64: Data-analysis templates will require higher accuracy, going
65: beyond adiabaticity; this remains a significant challenge.
66:
67: We describe an explicit computational procedure for obtaining
68: waveforms based on Mino's result, for the case of a point particle
69: coupled to a scalar field. We derive an explicit expression for the
70: time-averaged time derivative of the Carter constant, and verify that
71: the expression correctly predicts that circular orbits remain circular
72: while evolving under
73: the influence of radiation reaction.
74: The derivation uses detailed properties of mode expansions, Green's
75: functions and bound geodesic orbits in the Kerr spacetime, which we
76: review in detail.
77: This paper is about three quarters review and one quarter
78: new material. The intent is to
79: give a complete and self-contained treatment of scalar radiation
80: reaction in the Kerr spacetime, in a single unified notation, starting
81: with the Kerr metric, and ending with formulae for the
82: time evolution of all three constants of the motion that are sufficiently
83: explicit to be used immediately in a numerical code.
84:
85:
86:
87:
88: \end{abstract}
89: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
90: \maketitle
91: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
92:
93:
94: \section{Summary and overview}
95:
96:
97: \subsection{Gravitational radiation reaction of point particles}
98:
99: A key unsolved problem in general relativity is to compute
100: the gravitational radiation produced by a small object spiralling into
101: a much larger black hole. This problem is of direct observational relevance.
102: Inspirals of compact objects into intermediate
103: mass black holes ($M \sim 10^2 - 10^3 M_\odot)$
104: may be observed by LIGO and other ground based interferometers
105: \cite{2002ApJ...581..438M}; recent observations suggest the existence of
106: black holes in this mass range \cite{Miller:2003sc,Fiorito:2004qh}.
107: In addition, a key source for the space-based gravitational wave
108: detector LISA is the
109: final epoch of inspiral of a stellar-mass compact object into a
110: massive ($M \sim 10^6 M_\odot$) black hole at the center of a galaxy.
111: Gair et. al. \cite{Gair:2004iv} have estimated that LISA should
112: see over a thousand such inspiral events during its multi-year mission
113: lifetime, based on Monte Carlo simulations of the dynamics of stellar
114: cusps by Freitag \cite{2003ApJ...583L..21F}.
115:
116:
117:
118: Observations of these signals will have several major scientific
119: payoffs \cite{Cutler:2002me}:
120:
121: \begin{itemize}
122:
123: \item From the observed waveform, one can measure the mass
124: and spin of the central black hole with fractional accuracies of order
125: $10^{-4}$ \cite{Poisson:1996tc,Barack:2003fp}.
126: %Observing many
127: %events will therefore provide a census of the masses and spins of the
128: %massive central black holes in non-active galactic nuclei like M31 and
129: %M32.
130: The spin can provide useful information about the growth
131: history (mergers versus accretion) of the black hole \cite{Hughes:2002ei}.
132:
133: \item Likewise, one obtains a census of the inspiralling
134: objects' masses with precision $\sim 10^{-4}$, teaching us about the
135: stellar mass function and mass segregation in the central parsec of
136: galactic nuclei.
137:
138: \item The measured event rate will teach us about
139: dynamics in the central parsec of galaxies.
140:
141: \item The gravitational waves will be wonderful tools for probing the
142: nature of black holes in galactic nuclei, allowing us to
143: observationally map, for the first time, the spacetime geometry of a
144: black hole, and providing a high precision test of general relativity
145: in the strong field regime \cite{Ryan:1995wh,Ryan:1997hg}.
146:
147: \end{itemize}
148:
149:
150: To be in the LISA waveband, the mass $M$ of the black hole must be in
151: the range $10^5 \, M_\odot \alt M \alt 10^7 \, M_\odot$
152: \cite{Finn:2000sy}. The ratio between the mass $\mu \sim (1-100)
153: M_\odot$ of the compact object and $M$ is therefore in the range
154: $10^{-7} \alt \mu/M \alt 10^{-3}$.
155: %The smallness of this mass ratio
156: %%is the foundation for approximation methods used to compute the
157: %signal.
158: These systems spend the last year or so of
159: their lives in the very relativistic regime close to the black hole
160: horizon, where the post-Newtonian approximation has completely broken
161: down, and emit $\sim M/\mu \sim 10^5$ cycles of waveform \cite{Finn:2000sy}.
162:
163: Realizing the above science goals will require accurate theoretical models
164: (templates) of the gravitational waves. This is because the method of matched
165: filtering will be used both to detect the signals buried in the
166: detector noise, and to measure the parameters characterizing detected
167: signals. The accuracy requirement is roughly that the template should
168: gain or lose no more than $\sim 1$ cycle of phase
169: compared to the true waveform over the $\sim 10^5$ cycles of inspiral.
170: These sources must therefore be modeled with a fractional
171: accuracy $\sim 10^{-5}$. The past several years have seen a
172: significant research effort in the relativity community aimed
173: at providing these accurate templates.
174:
175:
176: To date, there have been several approaches to generating
177: waveforms. The foundation for all the approaches is the fact that, since
178: $\mu/M \ll 1$, the field of the compact object can be treated as a
179: linear perturbation to the large black hole's gravitational field. On
180: short timescales, the
181: compact object moves on a geodesic of the Kerr geometry,
182: characterized by its conserved energy $E$,
183: $z$-component of angular momentum $L_z$, and Carter
184: constant $Q$.
185: Over longer timescales, radiation reaction causes the parameters $E$,
186: $L_z$ and $Q$ to evolve and the orbit to shrink.
187:
188:
189: The various approaches are:
190:
191:
192: \begin{enumerate}
193:
194: \item {\it Use of post-Newtonian methods:}
195: Fairly crude waveforms can be obtained using post-Newtonian methods
196: \cite{Glampedakis:2002cb,Barack:2003fp}.
197: These have been used to approximately scope out LISA's
198: ability to detect inspiral events \cite{Gair:2004iv} and
199: to measure the waveform's parameters \cite{Barack:2003fp}.
200: However, since the orbital speeds are a substantial fraction of the speed
201: of light, these waveforms are insufficiently accurate for the eventual
202: detection and data analysis of real signals.
203:
204: \item {\it Use of conservation laws:}
205: In this approach
206: \cite{Cutler:1994pb,Shibata:1994xk,Glampedakis:2002ya,Hughes:1999bq}
207: one uses the Teukolsky-Sasaki-Nakamura (TSN) formalism
208: \cite{Teukolsky:1972my,Teukolsky:1973ha,Sasaki:1981sx} to compute
209: the fluxes
210: of energy $E$ and angular momentum $L_z$ to infinity and down the
211: black hole horizon generated by a
212: a compact object on a geodesic orbit.
213: Imposing global conservation of energy and
214: angular momentum, one infers the rates of change of the orbital energy
215: and angular momentum. For certain special classes of orbits (circular
216: and equatorial orbits) this
217: provides enough information that one can also infer the rate of change
218: of the Carter constant $Q$, and thus the inspiralling trajectory.
219:
220:
221:
222:
223: \item {\it Direct computation of the self-force:}
224: In this more fundamental approach one computes the self-force or
225: radiation-reaction force arising
226: from the interaction of the compact object with its own gravitational
227: field.
228: %This force is analogous to the
229: %Abraham-Lorentz-Dirac force that acts on accelerated electric charges.
230: A formal expression for this force in a general vacuum spacetime in terms
231: of the retarded Green's function was computed several years ago
232: \cite{Mino:1997nk,Quinn:1997am}.
233: Translating this expression into a practical computational
234: scheme for Kerr black holes is very difficult and is still in
235: progress. Roughly 100 papers
236: devoted to this problem have appeared in the last few years; see, for
237: example Poisson \cite{Poisson:2003nc}
238: for an overview and
239: references.
240:
241:
242:
243: \item {\it Time-domain numerical simulations:}
244: Another technique is to numerically integrate the
245: Teukolsky equation as a 2+1 PDE in the time domain
246: \cite{Krivan:1997hc,Burko:2002bt,Scheel:2003vs,Martel:2003,Lopez-Aleman:2003ik,Khanna:2003qv,Pazos-Avalos:2004rp},
247: and to model the compact object as a
248: finite-sized source. This approach faces considerable challenges:
249: (i) There is a separation of timescales --- the
250: orbital period is much shorter than the radiation reaction timescale.
251: (ii) There is a separation of lengthscales --- the compact object is
252: much smaller than the black hole. (iii) The self-field of the
253: small object must be computed with extremely high accuracy, as the
254: piece of the self-field responsible for the self-force is a tiny
255: fraction ($\sim \mu/M$) of the divergent self-field.
256: This approach may eventually be competitive with (ii) and (iii) but is
257: currently somewhat far from being competitive [flux accuracies are
258: $\sim 10\%$ as compared to $\sim 10^{-6}$ for (ii)].
259: \end{enumerate}
260:
261:
262:
263: \subsection{Radiation reaction in the adiabatic regime}
264:
265:
266: A key feature of these systems is that they evolve adiabatically: the
267: radiation reaction timescale $\sim M^2/\mu$ is much longer than the
268: orbital timescale $\sim M$, by a factor of the inverse $M/\mu$ of the
269: mass ratio. This has implications for the nature of the signal.
270: The self-acceleration ${\vec a}$ of the compact object can be expanded
271: in powers of the mass ratio as
272: \begin{equation}
273: {\vec a} = \frac{\mu}{M} \left[ {\vec a}_{1,{\rm diss}} + {\vec
274: a}_{1,{\rm cons}} + \frac{\mu}{M} {\vec a}_{2,{\rm diss}} +
275: \frac{\mu}{M} {\vec a}_{2,{\rm cons}} + O \left( \frac{\mu^2}{M^2}
276: \right) \right]\;.
277: \label{eq:selfaccel}
278: \end{equation}
279: Here ${\vec a}_{1,{\rm diss}}$ and ${\vec a}_{1,{\rm cons}}$ are the
280: dissipative and conservative pieces of the leading-order
281: self-acceleration computed in Refs.\ \cite{Mino:1997nk,Quinn:1997am}.
282: Similarly, ${\vec a}_{2,{\rm diss}}$ and ${\vec a}_{2,{\rm cons}}$ are
283: the corresponding pieces of the first correction to the
284: self-acceleration, which has not yet been computed (although see Ref.\
285: \cite{Rosenthal:2005ju} for work in this direction).
286:
287:
288: The effect of the dissipative pieces of the self force
289: will accumulate secularly, while the effect of the conservative pieces
290: will not. Hence the effect of the
291: dissipative pieces on the phase of the orbit will be larger than that of the
292: conservative pieces by a factor of the number of cycles of inspiral,
293: $\sim M/\mu$. Consider now, for example, the azimuthal phase
294: $\Phix(t)$ of the orbit. This can be
295: expanded in powers of the mass ratio using a two-timescale expansion
296: as \cite{dfh05,Hinderer}
297: %\footnote{Here we focus on the
298: % piece of the phase that grows linearly with time in the absence of
299: %radiation reaction. There are also oscillatory terms which are
300: %unimportant for this discussion; see Eq.\ (3.23) of Ref.\
301: % \cite{Drasco:2004tv}.}
302: \begin{equation}
303: \Phix(t) = \frac{M}{\mu} \left[ \Phix_1(t,t_1) + \frac{\mu}{M} \Phix_2(t,t_1) +
304: O\left( \frac{\mu^2}{M^2}\right) \right]\;,
305: \label{eq:emri_phase}
306: \end{equation}
307: where the leading-order phase
308: $\Phix_1(t,t_1)$ is determined by ${\vec a}_{1,{\rm diss}}$,
309: and the higher order correction $\Phix_2(t,t_1)$ to the phase is
310: determined by ${\vec a}_{1,{\rm cons}}$ and ${\vec a}_{2,{\rm diss}}$.
311: Here $t_1$ is a ``slow'' time variable which satisfies $dt_1/dt =
312: O(\mu/M)$.
313:
314: We shall call leading-order waveforms containing only terms
315: corresponding to
316: $\Phix_1(t,t_1)$ {\it adiabatic waveforms}. From Eq.\
317: (\ref{eq:emri_phase}) these waveforms are accurate
318: to $\sim 1$ cycle over the inspiral (more precisely, the phase error
319: is independent of $\mu/M$ in the limit $\mu/M \to 0$).
320: To compute adiabatic waveforms we need only keep the
321: dissipative piece of the self-force, and we can discard any
322: conservative pieces. The conservation-law method discussed above
323: yields waveforms that are accurate to this leading order, but is limited in that it
324: cannot be used for generic orbits.
325:
326:
327:
328: Adiabatic waveforms will likely be sufficiently accurate for detecting
329: inspiral events with LISA \cite{dfh05}. The correction $\Phix_2$ to
330: the phase of the waveform can be estimated using the post-Newtonian
331: approximation, and amounts to $\alt 1$ cycle over the
332: inspiral for typical parameter values.
333: While the post-Newtonian approximation is not strictly valid in the
334: highly relativistic regime near the horizon of interest here, it
335: suffices to give some indication of the accuracy of adiabatic
336: waveforms. This computation is described in \ref{sec:accuracy}.
337:
338:
339:
340: Recently Yasushi Mino derived a key result that paves the way for
341: computations of adiabatic waveforms for generic inspirals in Kerr
342: \cite{Mino:2003yg}.
343: Mino showed that the time average of the self-force formula of
344: \cite{Mino:1997nk,Quinn:1997am} for bound orbits in Kerr yields the
345: same result as the
346: gradient of one half the difference between the retarded and advanced
347: metric perturbations.
348: This ``half retarded minus half advanced'' prescription is the
349: standard result for electromagnetic radiation reaction in flat
350: spacetime due to Dirac \cite{Dirac:1938nz}.
351: This prescription was also posited, without proof, for scalar,
352: electromagnetic and gravitational radiation reaction in the Kerr
353: spacetime by Gal'tsov \cite{Galtsov:1982}.
354: However, prior to Mino's analysis, the prescription was not generally
355: thought to be applicable or relevant in Kerr.
356:
357: As one might expect, knowledge of the infinite-time-averaged self-force is
358: sufficient to compute the leading-order, adiabatic waveforms
359: \cite{Mino:2003yg}, although there are subtleties related to the
360: choice of gauge \cite{2005PThPh.113..733M,2005Minob,2005Minoc}. The
361: sufficiency can also be established using a two-timescale expansion
362: \cite{Hinderer}.
363:
364:
365: \subsection{Implementing Mino's prescription for generic inspirals}
366:
367:
368: In this paper we apply Mino's result to compute an explicit expression
369: for the time-averaged time derivative of the Carter constant. We
370: specialize to the case of a particle endowed with a scalar charge,
371: coupled to a scalar field. This computation is useful as a warm up
372: exercise for the more complicated case of a particle emitting
373: gravitational waves \cite{tensorcase}.
374:
375:
376: We start in Sec.\ \ref{sec:scalar} by describing the model of a point
377: particle coupled to a scalar field, and review how the self-force
378: causes both an acceleration of the particle's motion and also an
379: evolution of the renormalized rest mass of the particle \cite{Quinn:2000wa}.
380: In Sec.\ \ref{sec:geosesics} we review the properties of generic bound
381: geodesic orbits in the Kerr spacetime. A crucial result we use later
382: is that the $r$ and $\theta$ motions are periodic when expressed as
383: functions of a particular time parameter we call Mino time
384: \cite{Mino:2003yg}, and that
385: the $t$ and $\phi$ motions consist of linearly growing terms plus
386: terms that are periodic with the period of the $r$-motion, plus terms
387: that are periodic with the period of the $\theta$-motion
388: \cite{Drasco:2004tv}.
389:
390: Section \ref{sec:minoproof} reviews Mino's derivation of the
391: half-retarded-minus-half-advanced prescription for radiation reaction
392: in the adiabatic limit \cite{Mino:2003yg}. In Sec.\ \ref{sec:modes} we
393: discuss a convenient basis of modes for solutions of the scalar wave
394: equation in Kerr, namely the ``in'', ``out'', ``up'' and ``down''
395: modes used by Chrzanowski \cite{Chrzanowski:1975wv} and Gal'tsov
396: \cite{Galtsov:1982}. We also review several key properties of these
397: modes including symmetry relations and relations between the various
398: reflection and transmission coefficients that appear in the
399: definitions of the modes. Section \ref{sec:retarded} reviews the
400: standard derivation of the mode expansion of the retarded Green's function,
401: and Sec.\ \ref{sec:radiative} reviews the derivation of the mode
402: expansion of the radiative
403: Green's function given by Gal'tsov \cite{Galtsov:1982}, correcting
404: several typos in Gal'tsov which are detailed in \ref{sec:typos}.
405:
406: Next, in Sec.\ \ref{sec:harmonic} we turn to the source term in the
407: wave equation for the scalar field. We review the derivation of
408: Drasco and Hughes of the harmonic decomposition of this source term in
409: terms of a discrete sum over frequencies in which harmonics of three
410: different fundamental frequencies occur \cite{Drasco:2004tv}.
411: We derive expressions for the mode coefficients in the expansion of
412: the retarded field near future null infinity and near the future event
413: horizon. These coefficients are expressed as integrals over a torus in
414: phase space, which is filled ergodically by
415: the geodesic motion.
416:
417: Section \ref{sec:Edot} combines the results of the preceding sections
418: to derive expressions for the time-averaged rates of change of two of the
419: conserved quantities of geodesic motion, namely the energy $E$ and the angular
420: momentum $L_z$. The expressions are derived from the radiative self
421: force, following Gal'tsov \cite{Galtsov:1982}. Gal'tsov also shows that
422: identical expressions are obtained by using the fluxes of energy and
423: angular momentum to infinity and down the black hole horizon.
424: [This result has recently been independently derived for
425: circular, equatorial orbits in Ref.\ \protect{\cite{Gralla:2005et}}.]
426: We also show that the time-averaged rate of change of the renormalized
427: rest mass of the particle vanishes.
428:
429: In section \ref{sec:Kdot} we derive an expression for the
430: time-averaged rate of change of the Carter constant, using the
431: radiative self-force and the mode expansion of the radiative
432: Green's function. This expression [Eq.\ (\ref{eq:dKdt}) below] is the
433: main new result in this paper. It involves two new amplitudes that
434: are computed in terms of integrals over the torus in phase space,
435: just as for the amplitudes appearing in the energy and angular
436: momentum fluxes. Finally, in Sec.\ \ref{sec:circular} we show
437: that our result correctly predicts the known result that circular
438: orbits remain circular while evolving under the influence of radiation reaction.
439: This prediction serves as a check of our result.
440:
441:
442:
443: As apparent from the above summary, about 25 percent of this paper is
444: new material, and the remaining 75 percent is review.
445: The intent is to
446: give a complete and self-contained treatment of scalar radiation
447: reaction in the Kerr spacetime, in a single unified notation, starting
448: with the Kerr metric, and ending with formulae for the
449: evolution of all three constants of the motion that are sufficiently
450: explicit to be used immediately in a numerical code.
451:
452:
453:
454:
455: \section{The scalar field model}
456: \label{sec:scalar}
457:
458: We consider a point particle of scalar charge $q$
459: coupled to a scalar field $\Phi$. We denote by $\mu_0$ the bare
460: rest mass of the particle, to be distinguished from a renormalized
461: rest mass $\mu$ which will occur below.
462: The particle moves in a spacetime
463: with metric $g_{\alpha\beta}$ which is fixed; we neglect the gravitational
464: waves generated by the particle. The worldline of the particle is
465: $x^\alpha = z^\alpha(\tau)$, where $\tau$ is proper time.
466: The action is taken to be
467: \be
468: S = - \frac{1}{2} \int d^4 x \sqrt{-g} (\nabla \Phi)^2 - \int d\tau \left\{\mu_0 - q \Phi[z(\tau)] \right\}.
469: \label{eq:action}
470: \ee
471: Varying this action with respect to the worldline yields the equation
472: of motion
473: \begin{equation}
474: (\mu_0 - q \Phi) a^\alpha = q (g^{\alpha\beta} + u^\alpha u^\beta) \nabla_\beta \Phi,
475: \label{eq:acc00}
476: \end{equation}
477: where $u^\alpha$ is the 4-velocity and $a^\alpha$ is the 4-acceleration.
478: Following Poisson \cite{Poisson:2003nc}, we define the renormalized mass
479: $\mu$ by\footnote{If we endow the particle with an electric charge $e$
480: and add the electromagnetic coupling term $e \int d\tau u^\alpha
481: A_\alpha$ to the action (\protect{\ref{eq:action}}), the equation of motion
482: (\ref{eq:acc}) becomes
483: $\mu a^\alpha = q (g^{\alpha\beta} + u^\alpha u^\beta) \nabla_\beta
484: \Phi + e F^{\alpha\beta} u_\beta$, where $F^{\alpha\beta}$ is the
485: Faraday tensor. This shows that the renormalized mass
486: (\ref{eq:mudef}) is the mass that would be measured by coupling to
487: other fields.}
488: \be
489: \mu(\tau) = \mu_0 - q \Phi[z(\tau)].
490: \label{eq:mudef}
491: \ee
492: The equation of motion (\ref{eq:acc00}) can then be written as
493: \begin{equation}
494: a^\alpha = \frac{q}{\mu} (g^{\alpha\beta} + u^\alpha u^\beta) \nabla_\beta \Phi.
495: \label{eq:acc}
496: \end{equation}
497: It is also useful to rewrite the definition (\ref{eq:mudef}) of the
498: renormalized mass in terms of a differential equation for its evolution with
499: time:
500: \be
501: \frac{d \mu}{d\tau}(\tau) = - q u^\alpha \nabla_\alpha \Phi.
502: \label{eq:dmudt0}
503: \ee
504: The phenomenon of evolution of renormalized rest-mass is discussed
505: further in Refs.\ \cite{Burko:2002ge,Burko:2002gf,Haas:2004kw}.
506: Equations (\ref{eq:acc}) and (\ref{eq:dmudt0}) can also
507: be combined to give the expression
508: \be
509: f^\alpha = u^\beta \nabla_\beta ( \mu u^\alpha) = q \nabla^\alpha \Phi.
510: \label{eq:tsf}
511: \ee
512: for the total self force. The
513: components of $f^\alpha$ parallel to and perpendicular to the four
514: velocity yield the quantity $d\mu/d\tau$ and $\mu$ times the self-acceleration,
515: respectively.
516:
517:
518:
519:
520:
521: Varying the action (\ref{eq:action}) with respect to the field $\Phi$
522: gives the equation of motion
523: \be
524: \Box \Phi(x) = {\cal T}(x),
525: \label{eq:wave}
526: \ee
527: where the scalar source is
528: \be
529: {\cal T}(x) = - q \int_{-\infty}^\infty d \tau \delta^{(4)}[x,z(\tau)].
530: \label{eq:sourcedef}
531: \ee
532: Here
533: \be
534: \delta^{(4)}(x,x') = \delta^{(4)}(x - x')/\sqrt{-g}
535: \label{eq:deltafndef}
536: \ee
537: is the
538: generalized Dirac delta function, where $\delta^{(4)}(x) =
539: \delta(x^0) \delta(x^1) \delta(x^2) \delta(x^3)$.
540: We will assume that there is no
541: incoming scalar radiation, so that the physical solution of the wave
542: equation (\ref{eq:wave}) is the retarded solution
543: \be
544: \fl
545: \Phi_{\rm ret}(x) = \int d^4 x' \sqrt{-g(x')} G_{\rm ret}(x,x') {\cal
546: T}(x') = - q \int d\tau G_{\rm ret}[x,z(\tau)].
547: \label{eq:Phiret}
548: \ee
549: Here $G_{\rm ret}(x,x')$ is the retarded Green's function for the
550: scalar wave equation (\ref{eq:wave}).
551:
552:
553: The retarded field (\ref{eq:Phiret})
554: must of course be regularized before
555: being inserted into the expression (\ref{eq:acc}) for the
556: self-acceleration and into the expression (\ref{eq:dmudt0}) for $d\mu/d\tau$, or
557: else divergent results will be obtained. The appropriate regularization
558: prescription has been derived by Quinn \cite{Quinn:2000wa}.
559: The regularized self-acceleration at the point $z^\alpha(\tau)$
560: is \cite{Quinn:2000wa}
561: \be
562: \fl
563: a^\alpha(\tau) = \frac{q^2}{\mu} (g^{\alpha\beta} + u^\alpha
564: u^\beta ) \left[ \frac{1}{6} R_{\beta\gamma} u^\gamma -
565: \lim_{\varepsilon \to 0} \int_{-\infty}^{\tau -\varepsilon} d\tau'
566: \nabla_\beta G_{\rm ret}[z(\tau),z(\tau')] \right],
567: \label{eq:selfa1}
568: \ee
569: and the regularized expression for $d\mu/d\tau$ is
570: \be
571: \frac{d \mu}{d \tau}(\tau) = -\frac{q^2}{12} R + q^2
572: \lim_{\varepsilon \to 0} \int_{-\infty}^{\tau -\varepsilon} d\tau'
573: u^\alpha \nabla_\alpha G_{\rm ret}[z(\tau),z(\tau')].
574: \label{eq:dmudt1}
575: \ee
576:
577:
578: As discussed in the introduction, we specialize in this paper to bound
579: motion about a Kerr black hole of
580: mass $M$ with $M \gg \mu$. In this case the terms in Eqs.\
581: (\ref{eq:selfa1}) and (\ref{eq:dmudt1}) involving the Ricci tensor
582: $R_{\alpha\beta}$ vanish.
583: Also, as long as the orbit is not very close to the innermost stable orbit,
584: the evolution of the orbit is adiabatic: the orbital evolution
585: timescale $\sim M^2/\mu$ is much longer than the orbital timescale
586: $\sim M$. This adiabaticity allows a significant simplification of
587: the formulae (\ref{eq:selfa1}) and (\ref{eq:dmudt1})
588: for the self-acceleration and for the rate of change of mass
589: $d\mu/d\tau$, as shown by Mino
590: \cite{Mino:2003yg}.
591: %For gravitational radiation reaction, the regularized
592: %self-acceleration expression that corresponds to Eq.\ (\ref{eq:selfa1})
593: %was derived by Mino, Sasaki and Tanaka \cite{Mino:1997nk} and by
594: %Quinn and Wald \cite{Quinn:1997am}.
595: %Mino showed that in the adiabatic limit in Kerr, the
596: %the Mino-Sasaki-Tanaka-Quinn-Wald expression reduces
597: %to the gradient of one half the difference
598: %between the retarded and advanced metric perturbations \cite{Mino:2003yg}.
599: Namely, these formulae
600: %The result is that the expressions
601: %(\ref{eq:selfa1}) and (\ref{eq:dmudt1}) for the self-acceleration
602: %$a^\alpha$ and the rate of change of mass $d\mu/d\tau$
603: reduce to the
604: simple expressions (\ref{eq:acc}) and (\ref{eq:dmudt0}), but
605: with $\Phi$ replaced
606: by the
607: radiative field
608: \begin{equation}
609: \Phi_{\rm rad} = \frac{1}{2} ( \Phi_{\rm ret} - \Phi_{\rm adv}).
610: \end{equation}
611: Here $\Phi_{\rm adv}$ is the advanced solution of Eq.\
612: (\ref{eq:wave}).
613: In Sec.\ \ref{sec:minoproof} below we review Mino's argument,
614: specialized to the scalar case.
615:
616:
617: To summarize, the starting point for our analysis is the
618: self-acceleration expression
619: \be
620: a^\alpha = \frac{q}{\mu} (g^{\alpha\beta} + u^\alpha u^\beta)
621: \nabla_\beta \Phi_{\rm rad}.
622: \label{eq:selfacc0}
623: \ee
624: together with the expression for the evolution of rest mass
625: \be
626: \frac{d \mu}{d \tau}(\tau) = - q u^\alpha \nabla_\alpha \Phi_{\rm
627: rad}.
628: \label{eq:dmudt2}
629: \ee
630: Equations (\ref{eq:selfacc0}) and (\ref{eq:dmudt2}) are equivalent to
631: the equation
632: \be
633: f^\alpha = u^\beta \nabla_\beta ( \mu u^\alpha) = q \nabla^\alpha
634: \Phi_{\rm rad}.
635: \label{eq:tsf1}
636: \ee
637: for the total self-force.
638:
639:
640: \section{Generic bound geodesics in the Kerr spacetime}
641: \label{sec:geosesics}
642:
643:
644:
645:
646: This section summarizes the notation and results of Drasco \& Hughes
647: \cite{Drasco:2004tv}, and also some of the results of Mino \cite{Mino:2003yg}, for
648: generic bound geodesic orbits in Kerr.
649:
650:
651:
652:
653:
654: \subsection{The Kerr spacetime}
655: \label{sec:kerr}
656:
657: In Boyer-Lindquist coordinates $(t,r,\theta,\phi)$, the Kerr metric is
658: \begin{eqnarray}
659: \fl
660: ds^2 = - \left(1 - \frac{2 M r}{\Sigma}\right) dt^2 - \frac{4 a \sin^2 \theta M r
661: }{\Sigma} dt d\phi + ( \varpi^4 - \Delta a^2 \sin^2 \theta)
662: \frac{\sin^2 \theta}{\Sigma} d\phi^2 \nonumber \\
663: \lo + \Sigma d\theta^2 +
664: \frac{\Sigma}{\Delta} dr^2.
665: \label{eq:kerrmetric}
666: \end{eqnarray}
667: Here
668: \begin{eqnarray}
669: \Sigma & \equiv & r^2 + a^2 \cos^2 \theta, \\
670: \Delta & \equiv & r^2 + a^2 - 2 M r, \\
671: \varpi & \equiv & \sqrt{r^2 + a^2},
672: \end{eqnarray}
673: and $M$, $a$ are the black hole mass and spin parameter.
674: Throughout the rest of this paper we use units in which $M=1$, for
675: simplicity. The
676: square root of the determinant of the metric is
677: \be
678: \sqrt{-g} = \Sigma \sin \theta,
679: \label{eq:detg}
680: \ee
681: and the wave operator is given by
682: \begin{eqnarray}
683: \fl
684: \Sigma \Box \Phi = - \left[ \frac{\varpi^4}{\Delta} - a^2 \sin^2
685: \theta \right] \Phi_{,tt} - \frac{4 a r}{\Delta} \Phi_{,t\phi} +
686: \left( \frac{1}{\sin^2 \theta} - \frac{a^2}{\Delta} \right)
687: \Phi_{,\phi\phi} \nonumber \\
688: \lo + (\Delta \Phi_{,r})_{,r} + \frac{1}{\sin \theta}
689: (\sin \theta \Phi_{,\theta})_{,\theta}.
690: \label{eq:waveoperator}
691: \end{eqnarray}
692: The differential operator that appears on the right hand side of this
693: equation is the Teukolsky differential operator
694: \cite{Teukolsky:1973ha} for spin $s=0$.
695:
696:
697:
698: We will use later the Kinnersley null tetrad ${\vec l}$, ${\vec n}$,
699: ${\vec m}$, ${\vec m}^*$, which is given by
700: \be
701: {\vec l} = \frac{\varpi^2}{\Delta} \partial_t + \partial_r +
702: \frac{a}{\Delta} \partial_\phi,
703: \label{eq:vecldef}
704: \ee
705: \be
706: {\vec n} = \frac{\varpi^2}{2 \Sigma} \partial_t - \frac{\Delta}{2
707: \Sigma} \partial_r + \frac{a}{2 \Sigma} \partial_\phi,
708: \label{eq:vecndef}
709: \ee
710: and
711: \be
712: {\vec m} = \frac{1}{\sqrt{2}(r + i a \cos\theta)} \left( i a
713: \sin\theta \partial_t + \partial_\theta + \frac{i}{\sin\theta}
714: \partial_\phi \right).
715: \ee
716: The corresponding one-forms are
717: \be
718: {\bf l} = - dt + a \sin^2 \theta d\phi + \frac{\Sigma}{\Delta} dr,
719: \ee
720: \be
721: {\bf n} = - \frac{\Delta}{2 \Sigma} dt + \frac{a \Delta \sin^2
722: \theta}{2 \Sigma} d\phi - \frac{1}{2} dr,
723: \ee
724: and
725: \be
726: {\bf m} = \frac{1}{\sqrt{2}(r + i a \cos\theta)} \left( - i a
727: \sin\theta dt + \Sigma d\theta + i \varpi^2 \sin \theta d\phi \right).
728: \ee
729: The basis vectors obey the orthonormality relations ${\vec l} \cdot {\vec n} = -1$
730: and ${\vec m} \cdot {\vec m}^* = 1$ while all other inner products
731: vanish. The metric can be written in terms of the basis one-forms as
732: \be
733: g_{\alpha\beta} = -2 l_{(\alpha} n_{\beta)} + 2 m_{(\alpha} m_{\beta)}^*.
734: \label{eq:gabformula}
735: \ee
736:
737:
738:
739: \subsection{Constants of the motion}
740:
741:
742: We define the conserved energy per unit rest mass $\mu$
743: \be
744: E = - {\vec u} \cdot \frac{\partial}{\partial t},
745: \label{eq:Edef}
746: \ee
747: the conserved $z$-component of angular momentum divided by $\mu M$
748: \be
749: L_z = {\vec u} \cdot \frac{\partial}{\partial \phi},
750: \label{eq:Lzdef}
751: \ee
752: and Carter constant divided by $\mu^2 M^2$
753: \be
754: Q = u_\theta^2 - a^2 \cos^2 \theta E^2 + \cot^2 \theta L_z^2 + a^2
755: \cos^2 \theta.
756: \label{eq:Qdef}
757: \ee
758: [From now on we will for simplicity call these dimensionless quantities ``energy'',
759: ``angular momentum'' and ``Carter constant''.]
760: The geodesic equations can then be written in the form \cite{Drasco:2004tv}
761: \begin{eqnarray}
762: \fl
763: \left(\frac{dr}{d\lambda}\right)^2 = \left[E(r^2+a^2)
764: - a L_z\right]^2- \Delta\left[r^2 + (L_z - a E)^2 +
765: Q\right]
766: \equiv V_r(r)\;,
767: \label{eq:rdot}\\
768: %
769: \fl
770: \left(\frac{d\theta}{d\lambda}\right)^2 = Q - \cot^2\theta L_z^2
771: -a^2\cos^2\theta(1 - E^2)
772: \equiv V_\theta(\theta)\;,
773: \label{eq:thetadot}\\
774: %
775: \fl
776: \frac{d\phi}{d\lambda} =
777: \csc^2\theta L_z + aE\left(\frac{r^2+a^2}{\Delta} - 1\right) -
778: \frac{a^2L_z}{\Delta}
779: \equiv V_\phi(r,\theta)\;,
780: \label{eq:phidot}\\
781: %
782: \fl
783: \frac{dt}{d\lambda} =
784: E\left[\frac{(r^2+a^2)^2}{\Delta} - a^2\sin^2\theta\right] +
785: aL_z\left(1 - \frac{r^2+a^2}{\Delta}\right)
786: \equiv V_t(r,\theta)\;.
787: \label{eq:tdot}
788: \end{eqnarray}
789: Here $\lambda$ is the Mino time parameter \cite{Mino:2003yg}, related to
790: proper time $\tau$ by
791: \be
792: d\lambda = \frac{1}{\Sigma} d\tau.
793: \label{eq:Minotime}
794: \ee
795: Also these equations define the potentials $V_r(r)$,
796: $V_\theta(\theta)$, $V_\phi(r,\theta)$ and $V_t(r,\theta)$
797: \footnote{These quantities were denoted $R(r)$, $\Theta(\theta)$,
798: $\Phi(r,\theta)$ and $T(r,\theta)$ in Ref.\ \protect{\cite{Drasco:2004tv}}. We no
799: not use this notation here since it would clash with the functions
800: $R$, $\Theta$ and $\Phi$ defined in Eq.\ (\protect{\ref{eq:ansatz}}) below.}.
801:
802:
803:
804: Sometimes it will be convenient to use
805: instead of the Carter constant $Q$ the quantity
806: \be
807: K = Q + (L_z - a E)^2.
808: \ee
809: For convenience we will also call this quantity the ``Carter constant''.
810: In the Schwarzschild limit $Q$ and $K$ are given by $Q =
811: L_x^2 + L_y^2$ and $K = L_x^2 + L_y^2 + L_z^2$.
812: The quantity $K$ can be written as
813: \be
814: K = K^{\alpha\beta} u_\alpha u_\beta,
815: \label{eq:Kdef}
816: \ee
817: where $K^{\alpha\beta}$ is the Killing tensor
818: \be
819: K^{\alpha\beta} = 2 \Sigma m^{(\alpha} {\bar m}^{\beta)} - a^2 \cos^2
820: \theta g^{\alpha\beta}.
821: \ee
822: Using the identity (\ref{eq:gabformula}) this can also be written as
823: \be
824: K^{\alpha\beta} = 2 \Sigma l^{(\alpha} n^{\beta)} + r^2
825: g^{\alpha\beta}.
826: \label{eq:Kalphabeta}
827: \ee
828: Using the formulae (\ref{eq:vecldef}) and (\ref{eq:vecndef}) for the
829: null vectors ${\vec l}$ and ${\vec n}$ together with the definitions
830: (\ref{eq:Edef}) and (\ref{eq:Lzdef}) of $E$ and $L_z$, we obtain from
831: Eq.\ (\ref{eq:Kalphabeta}) the following formula for $K$:
832: \be
833: K = \frac{1}{\Delta} (\varpi^2 E - a L_z)^2 - \Delta u_r^2 - r^2.
834: \label{eq:Kformula1}
835: \ee
836: Solving this for $u_r$ gives
837: \be
838: u_r = \pm \sqrt{ \frac{1}{\Delta^2} ( \varpi^2 E - a L_z)^2 -
839: \frac{r^2}{\Delta} - \frac{K}{\Delta} };
840: \label{eq:urformula}
841: \ee
842: this formula will be useful later.
843:
844:
845: \subsection{Parameterization of solutions}
846: \label{sec:geodesicparameters}
847:
848: Following Mino \cite{Mino:2003yg}, we parameterize any geodesic by seven
849: parameters:
850: \be
851: E,L_z,Q,\lambda_{r0},\lambda_{\theta0},t_0,\phi_0.
852: \ee
853: Here $t_0$ and $\phi_0$ are the values of $t(\lambda)$ and
854: $\phi(\lambda)$ at $\lambda = 0$. The quantity $\lambda_{r0}$ is the
855: value of $\lambda$ nearest to $\lambda=0$ for which $r(\lambda) =
856: r_{\rm min}$, where $r_{\rm min}$ is the minimum value of $r$ attained
857: on the geodesic. Similarly $\lambda_{\theta0}$ is the
858: value of $\lambda$ nearest to $\lambda=0$ for which $\theta(\lambda) =
859: \theta_{\rm min}$, where $\theta_{\rm min}$ is the minimum value of
860: $\theta$ attained on the geodesic.
861: This parameterization is degenerate because of the freedom to
862: reparametrize the geodesic via $\lambda \to \lambda + \Delta
863: \lambda$. We discuss this degeneracy further in Sec.\
864: \ref{sec:degeneracy}.
865:
866: Frequently in this paper we will focus on the {\it fiducial
867: geodesic} associated with the constants $E$, $L_z$ and $Q$, namely
868: the geodesic with
869: \be
870: \lambda_{r0} = \lambda_{\theta0} = t_0 = \phi_0 =0.
871: \ee
872:
873:
874: \subsection{Motions in $r$ and $\theta$}
875:
876:
877:
878:
879: It follows from the geodesic equations (\ref{eq:rdot}) and
880: (\ref{eq:thetadot}) that the functions
881: $r(\lambda)$ and $\theta(\lambda)$ are periodic. We denote the
882: periods by $\Lambda_r$ and $\Lambda_\theta$, respectively, so
883: \be
884: r(\lambda + \Lambda_r) = r(\lambda), \ \ \ \ \ \
885: \theta(\lambda + \Lambda_\theta) = \theta(\lambda).
886: \label{eq:periodic}
887: \ee
888: Using the initial condition $r(\lambda_{r0}) = r_{\rm min}$ we can
889: write the solution $r(\lambda)$ to Eq.\ (\ref{eq:rdot}) explicitly as
890: \be
891: r(\lambda) = {\hat r}(\lambda - \lambda_{r0}),
892: \label{eq:rmotion}
893: \ee
894: where the function ${\hat r}(\lambda)$ is defined by
895: \be
896: \int_{r_{\rm min}}^{{\hat r}(\lambda)} \frac{dr}{\pm \sqrt{V_r(r)}} =
897: \lambda.
898: \label{eq:hatrdef}
899: \ee
900: Similarly we can
901: write the solution $\theta(\lambda)$ to Eq.\ (\ref{eq:thetadot}) explicitly as
902: \be
903: \theta(\lambda) = {\hat \theta}(\lambda - \lambda_{\theta0}),
904: \label{eq:thetamotion}
905: \ee
906: where the function ${\hat \theta}(\lambda)$ is defined by
907: \be
908: \int_{\theta_{\rm min}}^{{\hat \theta}(\lambda)} \frac{d\theta}{\pm \sqrt{V_\theta(\theta)}} = \lambda.
909: \label{eq:hatthetadef}
910: \ee
911: The functions ${\hat r}(\lambda)$ and ${\hat \theta}(\lambda)$ are
912: just the $r$ and $\theta$ motions for the fiducial geodesic.
913:
914:
915: \subsection{Motion in $t$}
916:
917: Next, the function $V_t(r,\theta)$ that appears on the right hand side
918: of Eq.\ (\ref{eq:tdot}) is a sum of a function of $r$ and a function
919: of $\theta$:
920: \be
921: V_t(r,\theta) = V_{tr}(r) + V_{t\theta}(\theta),
922: \ee
923: where
924: $V_{tr}(r) = E \varpi^4/\Delta + aL_z(1 - \varpi^2/\Delta)$ and
925: $V_{t\theta}(\theta) = - a^2 E \sin^2 \theta$.
926: Therefore using $t(0) = t_0$ we obtain
927: \be
928: t(\lambda) = t_0 + \int_0^\lambda d\lambda' \left\{ V_{tr}[r(\lambda')] +
929: V_{t\theta}[\theta(\lambda')] \right\}.
930: \label{eq:tmotion0}
931: \ee
932: Next we define the averaged value $\langle V_{tr} \rangle$ of $V_{tr}$ to be
933: \be
934: \langle V_{tr} \rangle = \frac{1}{\Lambda_r} \int_0^{\Lambda_r} d\lambda
935: \
936: V_{tr}[r(\lambda)] = \frac{1}{\Lambda_r} \int_0^{\Lambda_r} d\lambda\
937: V_{tr}[{\hat r}(\lambda)].
938: \ee
939: Here the second equality follows from the representation (\ref{eq:rmotion}) of
940: the $r$ motion together with the periodicity condition (\ref{eq:periodic}).
941: Similarly we define
942: \be
943: \langle V_{t\theta} \rangle = \frac{1}{\Lambda_\theta} \int_0^{\Lambda_\theta} d\lambda
944: \
945: V_{t\theta}[\theta(\lambda)] = \frac{1}{\Lambda_\theta} \int_0^{\Lambda_\theta} d\lambda\
946: V_{t\theta}[{\hat \theta}(\lambda)].
947: \ee
948: Inserting these definitions into Eq.\ (\ref{eq:tmotion0})
949: allows us to write $t(\lambda)$ as a sum of a linear term and terms
950: that are periodic \cite{Drasco:2004tv}:
951: \begin{eqnarray}
952: \label{eq:tmotion00}
953: t(\lambda) &=& t_0 + \Gamma \lambda + \Delta t_r(\lambda) + \Delta
954: t_\theta(\lambda) \\
955: &\equiv& t_0 + \Gamma \lambda + \Delta t(\lambda).
956: \label{eq:tmotion1}
957: \end{eqnarray}
958: Here we have defined the constant
959: \be
960: \Gamma = \langle V_{tr} \rangle + \langle V_{t\theta} \rangle,
961: \label{eq:Gammadef}
962: \ee
963: and the functions
964: \be
965: \Delta t_r(\lambda) = \int_0^\lambda d\lambda' \bigg\{
966: V_{tr}[r(\lambda')] - \langle V_{tr} \rangle \bigg\},
967: \label{eq:deltatrdef}
968: \ee
969: \be
970: \Delta t_\theta(\lambda) = \int_0^\lambda d\lambda' \bigg\{
971: V_{t\theta}[\theta(\lambda')] - \langle V_{t\theta} \rangle \bigg\}.
972: \label{eq:deltatthetadef}
973: \ee
974: The key property of these functions is that they are periodic:
975: \be
976: \Delta t_r(\lambda + \Lambda_r) = \Delta t_r(\lambda),
977: \ee
978: \be
979: \Delta t_\theta(\lambda + \Lambda_\theta) = \Delta t_\theta(\lambda);
980: \ee
981: this follows from the definitions (\ref{eq:deltatrdef}) and
982: (\ref{eq:deltatthetadef}) together with the periodicity condition
983: (\ref{eq:periodic}). We can exhibit the dependence of these
984: functions on the parameters $\lambda_{r0}$ and $\lambda_{\theta0}$
985: by substituting the formulae (\ref{eq:rmotion}) and
986: (\ref{eq:thetamotion}) for $r(\lambda)$ and $\theta(\lambda)$ into
987: Eqs.\ (\ref{eq:deltatrdef}) and
988: (\ref{eq:deltatthetadef}). The result is
989: \be
990: \Delta t_r(\lambda) = {\hat t}_r(\lambda - \lambda_{r0}) - {\hat
991: t}_r(-\lambda_{r0}),
992: \label{eq:deltatrformula}
993: \ee
994: \be
995: \Delta t_\theta(\lambda) = {\hat t}_\theta(\lambda - \lambda_{\theta0}) - {\hat
996: t}_\theta(-\lambda_{\theta0}),
997: \label{eq:deltatthetaformula}
998: \ee
999: where the functions ${\hat t}_r(\lambda)$ and ${\hat t}_\theta(\lambda)$ are defined by
1000: \be
1001: {\hat t}_r(\lambda) = \int_0^\lambda d\lambda' \bigg\{
1002: V_{tr}[{\hat r}(\lambda')] - \langle V_{tr} \rangle \bigg\},
1003: \label{eq:hattrdef}
1004: \ee
1005: \be
1006: {\hat t}_\theta(\lambda) = \int_0^\lambda d\lambda' \bigg\{
1007: V_{t\theta}[{\hat \theta}(\lambda')] - \langle V_{t\theta} \rangle \bigg\}.
1008: \label{eq:hattthetadef}
1009: \ee
1010:
1011:
1012: \subsection{Motion in $\phi$}
1013:
1014: The motion in $\phi$ can be analyzed in exactly the same way as the
1015: motion in $t$. First, the function $V_\phi(r,\theta)$ that appears on
1016: the right hand side of Eq.\ (\ref{eq:phidot}) is a sum of a function
1017: of $r$ and a function of $\theta$:
1018: \be
1019: V_\phi(r,\theta) = V_{\phi r}(r) + V_{\phi\theta}(\theta),
1020: \ee
1021: where
1022: $V_{\phi r}(r)=a E (\varpi^2/\Delta - 1) - a^2 L_z/\Delta$ and
1023: $V_{\phi\theta}(\theta) = \csc^2\theta L_z$. Therefore using $\phi(0)
1024: = \phi_0$ we obtain
1025: \be
1026: \phi(\lambda) = \phi_0 + \int_0^\lambda d\lambda' \left\{ V_{\phi r}[r(\lambda')] +
1027: V_{\phi\theta}[\theta(\lambda')] \right\}.
1028: \label{eq:phimotion0}
1029: \ee
1030: Next we define the averaged value $\langle V_{\phi r} \rangle$ of
1031: $V_{\phi r}$ to be
1032: \be
1033: \langle V_{\phi r} \rangle = \frac{1}{\Lambda_r} \int_0^{\Lambda_r} d\lambda
1034: \
1035: V_{\phi r}[r(\lambda)] = \frac{1}{\Lambda_r} \int_0^{\Lambda_r} d\lambda\
1036: V_{\phi r}[{\hat r}(\lambda)].
1037: \ee
1038: Here the second equality follows from the representation (\ref{eq:rmotion}) of
1039: the $r$ motion together with the periodicity condition (\ref{eq:periodic}).
1040: Similarly we define
1041: \be
1042: \langle V_{\phi\theta} \rangle = \frac{1}{\Lambda_\theta}
1043: \int_0^{\Lambda_\theta} d\lambda
1044: \
1045: V_{\phi\theta}[\theta(\lambda)] = \frac{1}{\Lambda_\theta} \int_0^{\Lambda_\theta} d\lambda\
1046: V_{\phi\theta}[{\hat \theta}(\lambda)].
1047: \ee
1048: Inserting these into Eq.\ (\ref{eq:phimotion0}) and using $\phi(0) = \phi_0$
1049: allows us to write $\phi(\lambda)$ as a sum of a linear term and terms
1050: that are periodic \cite{Drasco:2004tv}:
1051: \begin{eqnarray}
1052: \label{eq:phimotion00}
1053: \phi(\lambda) &=& \phi_0 + \Upsilon_\phi \lambda + \Delta \phi_r(\lambda) + \Delta
1054: \phi_\theta(\lambda) \\
1055: &\equiv& \phi_0 + \Upsilon_\phi \lambda + \Delta \phi(\lambda).
1056: \label{eq:phimotion1}
1057: \end{eqnarray}
1058: Here we have defined the constant
1059: \be
1060: \Upsilon_\phi = \langle V_{\phi r} \rangle + \langle V_{\phi\theta} \rangle,
1061: \label{eq:Upsilonphidef}
1062: \ee
1063: and the functions
1064: \be
1065: \Delta \phi_r(\lambda) = \int_0^\lambda d\lambda' \bigg\{
1066: V_{\phi r}[r(\lambda')] - \langle V_{\phi r} \rangle \bigg\},
1067: \label{eq:deltaphirdef}
1068: \ee
1069: \be
1070: \Delta \phi_\theta(\lambda) = \int_0^\lambda d\lambda' \bigg\{
1071: V_{\phi\theta}[\theta(\lambda')] - \langle V_{\phi\theta} \rangle \bigg\}.
1072: \label{eq:deltaphithetadef}
1073: \ee
1074: The key property of these functions is that they are periodic:
1075: \be
1076: \Delta \phi_r(\lambda + \Lambda_r) = \Delta \phi_r(\lambda),
1077: \ee
1078: \be
1079: \Delta \phi_\theta(\lambda + \Lambda_\theta) = \Delta \phi_\theta(\lambda);
1080: \ee
1081: this follows from the definitions (\ref{eq:deltaphirdef}) and
1082: (\ref{eq:deltaphithetadef}) together with the periodicity condition
1083: (\ref{eq:periodic}). We can exhibit the dependence of these
1084: functions on the parameters $\lambda_{r0}$ and $\lambda_{\theta0}$
1085: by substituting the formulae (\ref{eq:rmotion}) and
1086: (\ref{eq:thetamotion}) for $r(\lambda)$ and $\theta(\lambda)$ into
1087: Eqs.\ (\ref{eq:deltaphirdef}) and
1088: (\ref{eq:deltaphithetadef}). The result is
1089: \be
1090: \Delta \phi_r(\lambda) = {\hat \phi}_r(\lambda - \lambda_{r0}) - {\hat
1091: \phi}_r(-\lambda_{r0}),
1092: \label{eq:deltaphirformula}
1093: \ee
1094: \be
1095: \Delta \phi_\theta(\lambda) = {\hat \phi}_\theta(\lambda -
1096: \lambda_{\theta0}) - {\hat \phi}_\theta(-\lambda_{\theta0}),
1097: \label{eq:deltaphithetaformula}
1098: \ee
1099: where the functions ${\hat \phi}_r$ and ${\hat \phi}_\theta$ are defined by
1100: \be
1101: {\hat \phi}_r(\lambda) = \int_0^\lambda d\lambda' \bigg\{
1102: V_{\phi r}[{\hat r}(\lambda')] - \langle V_{\phi r} \rangle \bigg\},
1103: \label{eq:hatphirdef}
1104: \ee
1105: \be
1106: {\hat \phi}_\theta(\lambda) = \int_0^\lambda d\lambda' \bigg\{
1107: V_{\phi\theta}[{\hat \theta}(\lambda')] - \langle V_{\phi\theta}
1108: \rangle \bigg\}.
1109: \label{eq:hatphithetadef}
1110: \ee
1111:
1112: \subsection{Re-parameterization freedom}
1113: \label{sec:degeneracy}
1114:
1115: Not all of the parameters $E$, $L_z$, $Q$, $\lambda_{r0}$,
1116: $\lambda_{\theta0}$, $t_0$ and $\phi_0$ that characterize the geodesic
1117: are independent. This is because of the freedom to change the
1118: dependent variable $\lambda$ via $\lambda \to {\tilde \lambda} = \lambda +
1119: \Delta \lambda$. Under this change of variable the parameters
1120: $\lambda_{r0}$ and $\lambda_{\theta0}$
1121: transform as
1122: \be
1123: \lambda_{r0} \to {\tilde \lambda}_{r0} = \lambda_{r0} + \Delta
1124: \lambda,
1125: \label{eq:t1}
1126: \ee
1127: \be
1128: \lambda_{\theta0} \to {\tilde \lambda}_{\theta0} = \lambda_{\theta0} +
1129: \Delta \lambda.
1130: \label{eq:t2}
1131: \ee
1132: We can compute how the parameters $t_0$ and $\phi_0$ transform as
1133: follows. Combining Eqs.\ (\ref{eq:tmotion1}),
1134: (\ref{eq:deltatrformula}) and (\ref{eq:deltatthetaformula})
1135: gives the following formula for the $t$ motion:
1136: \be
1137: \fl
1138: t = t_0 + \Gamma \lambda
1139: + {\hat t}_r(\lambda - \lambda_{r0}) - {\hat t}_r(-\lambda_{r0})
1140: + {\hat t}_\theta(\lambda - \lambda_{\theta0}) - {\hat
1141: t}_\theta(-\lambda_{\theta0}).
1142: \ee
1143: Rewriting the right hand side in terms of ${\tilde \lambda}$, ${\tilde \lambda}_{r0}$
1144: and ${\tilde \lambda}_{\theta0}$ yields
1145: \begin{eqnarray}
1146: \fl
1147: t \ = t_0 + \Gamma {\tilde \lambda} - \Gamma \Delta \lambda
1148: + {\hat t}_r({\tilde \lambda} - {\tilde \lambda}_{r0}) - {\hat t}_r(-\lambda_{r0})
1149: + {\hat t}_\theta({\tilde \lambda} - {\tilde \lambda}_{\theta0}) - {\hat
1150: t}_\theta(-\lambda_{\theta0}) \nn \\
1151: \fl
1152: \ \ \ \equiv {\tilde t}_0 + \Gamma {\tilde \lambda}
1153: + {\hat t}_r({\tilde \lambda} - {\tilde \lambda}_{r0}) - {\hat
1154: t}_r(-{\tilde \lambda}_{r0})
1155: + {\hat t}_\theta({\tilde \lambda} - {\tilde \lambda}_{\theta0}) - {\hat
1156: t}_\theta(-{\tilde \lambda}_{\theta0}).
1157: \end{eqnarray}
1158: Comparing the first and second lines here allows us to read off the
1159: value of ${\tilde t}_0$:
1160: \be
1161: \fl
1162: {\tilde t}_0 = t_0 - \Gamma \Delta \lambda
1163: + {\hat t}_r(-\lambda_{r0} - \Delta \lambda) - {\hat
1164: t}_r(-\lambda_{r0})
1165: + {\hat t}_\theta(-\lambda_{\theta0} - \Delta \lambda) - {\hat t}_\theta(-\lambda_{\theta0}).
1166: \label{eq:t3}
1167: \ee
1168: Similarly we obtain
1169: \be
1170: \fl
1171: \phi_0 \to {\tilde \phi}_0 = \phi_0 - \Upsilon_\phi \Delta \lambda
1172: + {\hat \phi}_r(-\lambda_{r0} - \Delta \lambda) - {\hat
1173: \phi}_r(-\lambda_{r0})
1174: + {\hat \phi}_\theta(-\lambda_{\theta0} - \Delta \lambda) - {\hat
1175: \phi}_\theta(-\lambda_{\theta0}).
1176: \label{eq:t4}
1177: \ee
1178: In Sec.\ \ref{sec:dep} below we will explicitly show that all of the amplitudes
1179: and waveforms we compute are invariant under these transformations.
1180:
1181:
1182: \subsection{Averages over Mino time in terms of angular variables}
1183:
1184: We will often encounter functions $F_\theta(\lambda)$ of Mino time
1185: $\lambda$ that are periodic with period $\Lambda_\theta$. For such
1186: functions the average over Mino time is given
1187: by
1188: \be
1189: \langle F_\theta \rangle_\lambda = \frac{1}{\Lambda_\theta} \int_0^{\Lambda_\theta}
1190: F_\theta(\lambda) d \lambda.
1191: \label{eq:thetaaverage}
1192: \ee
1193: In order to compute such averages it will be convenient to
1194: change the variable of integration from $\lambda$ to a new variable
1195: $\chi$.
1196: This variable is a generalization of the Newtonian true anomaly and is defined as
1197: follows \cite{Wilkins:1972rs,Hughes:1999bq}.
1198: The potential $V_\theta(\theta)$
1199: defined by Eq.\ (\ref{eq:thetadot}) can be written in terms of the
1200: variable
1201: \be
1202: z \equiv \cos^2 \theta
1203: \ee
1204: as
1205: \be
1206: V_\theta(z) = \frac{\beta (z - z_+) (z - z_-)}{1 - z}.
1207: \label{eq:Thetaformula}
1208: \ee
1209: Here $\beta = a^2(1-E^2)$ and $z_-$ and $z_+$ are the two zeros of
1210: $V_\theta(z)$. They are
1211: ordered such that $0 \le z_- \le 1 \le z_+$. The variable $\chi$ is
1212: defined by
1213: \be
1214: \cos^2 {\hat \theta}(\lambda) = z_- \cos^2 \chi,
1215: \ee
1216: together with the requirement that $\chi$ increases monotonically as
1217: $\lambda$ increases.
1218: From the definition (\ref{eq:hatthetadef}) of ${\hat \theta}(\lambda)$
1219: together with the formula (\ref{eq:Thetaformula}) for $V_\theta(z)$ we
1220: get
1221: \be
1222: \frac{d\chi}{d\lambda} = \sqrt{\beta (z_+ - z_- \cos^2 \chi)}.
1223: \ee
1224: Therefore the average (\ref{eq:thetaaverage}) can be written as
1225: \be
1226: \langle F_\theta \rangle_\lambda = \frac{1}{\Lambda_\theta} \int_0^{2
1227: \pi} d\chi
1228: \frac{F_\theta[\lambda(\chi)] }{\sqrt{\beta(z_+ - z_- \cos^2 \chi)}}.
1229: \label{eq:thetaaverage1}
1230: \ee
1231: %Using $\langle 1 \rangle = 1$ gives an expression for
1232: %$\Lambda_\theta$:
1233: %\be
1234: %\Lambda_\theta = \frac{4}{\sqrt{\beta z_+}} \int_0^{\pi/2}
1235: %\frac{1}{\sqrt{1 - (z_-/z_+) \cos^2 \chi}} = \frac{4}{\sqrt{\beta z_+}}
1236: % K(\sqrt{z_-/z_+}),
1237: %\ee
1238: %where $K$ is the complete elliptic integral of the first kind. Using
1239: %this in Eq.\ (\ref{eq:thetaaverage1}) gives
1240: %\be
1241: %\langle F_\theta \rangle_\lambda = \frac{1}{4 K(\sqrt{z_-/z_+})} \int_0^{2
1242: % \pi} d\chi
1243: %\frac{F_\theta[\lambda(\chi)] }{\sqrt{1 - (z_-/z_+) \cos^2 \chi}}.
1244: %\label{eq:thetaaverage2}
1245: %\ee
1246:
1247:
1248: A similar analysis applies to averages of functions $F_r(\lambda)$
1249: that are periodic with period $\Lambda_r$ \cite{Drasco:2004tv}. The
1250: average over $\lambda$ of such a function is
1251: \be
1252: \langle F_r \rangle_\lambda = \frac{1}{\Lambda_r} \int_0^{\Lambda_r}
1253: F_r(\lambda) d \lambda.
1254: \label{eq:raverage}
1255: \ee
1256: Now the
1257: potential $V_r(r)$ in
1258: Eq.\ (\ref{eq:rdot}) can be written as \cite{Drasco:2004tv}
1259: \be
1260: V_r(r) = (1-E^2)(r_1-r)(r-r_2)(r-r_3)(r-r_4),
1261: \ee
1262: where $r_1$, $r_2$, $r_3$ and $r_4$ are the four roots of the quartic
1263: $V_r(r)=0$, ordered such that $r_4 \le r_3 \le r_2 \le r_1$.
1264: For stable orbits the motion takes place in $r_2 \le r \le r_1$.
1265: The orbital eccentricity $\varepsilon$ and semi-latus rectum $p$ are
1266: defined by
1267: \be
1268: r_1 = \frac{p}{1-\varepsilon}
1269: \ee
1270: and
1271: \be
1272: r_2 = \frac{p}{1 + \varepsilon}.
1273: \ee
1274: We also define the parameters $p_3$ and $p_4$ by
1275: \be
1276: r_3 = \frac{p_3 }{1 - \varepsilon}, \ \ \ \ \
1277: r_4 = \frac{p_4 }{1 + \varepsilon}.
1278: \ee
1279: Then following Ref.\ \cite{Cutler:1994pb}
1280: we define the parameter $\psi$ by
1281: \be
1282: {\hat r}(\lambda) = \frac{p }{1 + \varepsilon \cos \psi}.
1283: \label{eq:psidef}
1284: \ee
1285: It follows from these definitions that \cite{Drasco:2004tv}
1286: \be
1287: \frac{d \psi}{d \lambda} = {\cal P}(\psi),
1288: \ee
1289: where
1290: \be
1291: \fl
1292: {\cal P}(\psi) = \frac{\sqrt{1-E^2}}{1 - \varepsilon^2}
1293: \left[ p(1-\varepsilon) - p_3(1 + \varepsilon \cos \psi) \right]^{1/2}
1294: \left[ p(1+\varepsilon) - p_4(1 + \varepsilon \cos \psi) \right]^{1/2}.
1295: \ee
1296: The average (\ref{eq:raverage}) over $\lambda$ is therefore
1297: \be
1298: \langle F_r \rangle_\lambda = \frac{1}{\Lambda_r} \int_0^{2 \pi} d\psi
1299: \frac{F_r(\lambda)}{{\cal P}(\psi)} = \frac{
1300: \int_0^{2 \pi} d\psi F_r(\lambda)/ {\cal P}(\psi)}
1301: {\int_0^{2 \pi} d\psi / {\cal P}(\psi)}.
1302: \label{eq:raverage1}
1303: \ee
1304:
1305:
1306: Note that the viewpoint on the new variables $\chi$,$\psi$ adopted
1307: here is slightly different to that in \cite{Drasco:2004tv}: Here they are
1308: defined in terms of the fiducial motions ${\hat r}(\lambda)$ and
1309: ${\hat \theta}(\lambda)$ instead of the actual motions $r(\lambda)$
1310: and $\theta(\lambda)$. The new viewpoint facilitates the computation
1311: of amplitudes and fluxes for arbitrary geodesics; if one is working
1312: with the fiducial geodesic the distinction is unimportant.
1313:
1314:
1315: \subsubsection{Bi-periodic functions}
1316:
1317: Finally suppose we have a function $F(\lambda_r,\lambda_\theta)$ of
1318: two parameters $\lambda_r$ and $\lambda_\theta$ which is biperiodic:
1319: \begin{eqnarray}
1320: F(\lambda_r + \Lambda_r, \lambda_\theta) = F(\lambda_r,\lambda_\theta)
1321: \nn \\
1322: F(\lambda_r , \lambda_\theta + \Lambda_\theta) =
1323: F(\lambda_r,\lambda_\theta).
1324: \end{eqnarray}
1325: The average value of this function is
1326: \be
1327: \langle F \rangle_\lambda = \frac{1}{\Lambda_r \Lambda_\theta}
1328: \int_0^{\Lambda_r} d\lambda_r \, \int_0^{\Lambda_\theta}
1329: d\lambda_\theta F(\lambda_r,\lambda_\theta).
1330: \ee
1331: By combining the results (\ref{eq:thetaaverage1}) and
1332: (\ref{eq:raverage1}) we can write this average as a double integral
1333: over the new variables $\chi$ and $\psi$:
1334: \be
1335: \langle F \rangle_\lambda = \frac{1}{\Lambda_r \Lambda_\theta}
1336: \int_0^{2 \pi} d\chi \int_0^{2 \pi} d\psi
1337: \frac{F[\lambda_r(\psi),\lambda_\theta(\chi)]}{ \sqrt{ \beta( z_+ -
1338: z_- \cos^2 \chi) } {\cal P}(\psi)}.
1339: \label{eq:averageidentity}
1340: \ee
1341: This formula will be used in later sections.
1342:
1343:
1344: \section{Mino's derivation of half retarded minus half advanced prescription}
1345: \label{sec:minoproof}
1346:
1347: In this section we review the proof by Mino \cite{Mino:2003yg}
1348: that for computing radiation reaction in the adiabatic limit in Kerr,
1349: one can use the ``half retarded minus half advanced'' prescription.
1350: We specialize to the scalar case.
1351:
1352: We start by defining some notation for self-forces.
1353: Suppose that we have a particle with scalar charge $q$ at a point ${\cal P}$ in a
1354: spacetime $M$ with metric $g_{\alpha\beta}$. Suppose that the 4-velocity
1355: of the particle at ${\cal P}$ is $u^\alpha$, and that we are given a
1356: solution $\Phi$ (not necessarily the retarded solution) of the wave
1357: equation (\ref{eq:wave}) for the scalar field, for which the source is
1358: a delta function on the geodesic determined by ${\cal P}$ and
1359: $u^\alpha$. The self-force on the particle is then some
1360: functional of ${\cal P}$, $u^\alpha$, $g_{\alpha\beta}$ and $\Phi$,
1361: which we write as
1362: \be
1363: f^\alpha\left[ {\cal P}, u^\alpha, g_{\alpha\beta}, \Phi \right].
1364: \ee
1365: Here we suppress the trivial dependence on $q$.
1366: Note that this functional does not depend on a choice of time
1367: orientation for the manifold, and also it is invariant under $u^\alpha
1368: \to - u^\alpha$.
1369:
1370: Next, we define the retarded self-force as
1371: \be
1372: f^\alpha_{\rm ret}\left[ {\cal P}, u^\alpha, g_{\alpha\beta} \right]
1373: = f^\alpha\left[ {\cal P}, u^\alpha, g_{\alpha\beta},
1374: \Phi_{\rm ret} \right],
1375: \label{eq:selfret}
1376: \ee
1377: where $\Phi_{\rm ret}$ is the retarded solution to the wave equation
1378: (\ref{eq:wave}) using the time orientation that is determined by
1379: demanding that $u^\alpha$ be future directed. Similarly we define the
1380: advanced self force by
1381: \be
1382: f^\alpha_{\rm adv}\left[ {\cal P}, u^\alpha, g_{\alpha\beta} \right]
1383: = f^\alpha\left[ {\cal P}, u^\alpha, g_{\alpha\beta},
1384: \Phi_{\rm adv} \right],
1385: \ee
1386: where $\Phi_{\rm adv}$ is the advanced solution.
1387: It follows from these definitions that
1388: \be
1389: f^\alpha_{\rm ret}\left[ {\cal P}, -u^\alpha, g_{\alpha\beta} \right]
1390: = f^\alpha_{\rm adv}\left[ {\cal P}, u^\alpha, g_{\alpha\beta}
1391: \right].
1392: \label{eq:flip}
1393: \ee
1394:
1395:
1396: \subsection{Properties of the self-force}
1397:
1398: The derivation of the half-retarded-minus-half-advanced prescription
1399: in the adiabatic limit
1400: rests on two properties of the self-force. The first
1401: property is the fact that the self-force can be computed by
1402: subtracting from the divergent field $\Phi$ a locally constructed
1403: singular field $\Phi_{\rm sing}$ that depends only on the spacetime geometry in
1404: the vicinity of the point ${\cal P}$ and on the four velocity at
1405: ${\cal P}$. This property is implicit in the work of Quinn \cite{Quinn:2000wa}
1406: and Quinn and Wald \cite{Quinn:1997am}, and was proved explicitly by
1407: Detweiler and Whiting \cite{Detweiler:2002mi}. We can write this
1408: property as
1409: \be
1410: f^\alpha\left[ {\cal P}, u^\alpha, g_{\alpha\beta}, \Phi \right] = q
1411: \nabla^\alpha \left( \Phi - \Phi_{\rm sing}[{\cal
1412: P},u^\alpha,g_{\alpha\beta}] \right).
1413: \label{eq:property0}
1414: \ee
1415:
1416: The second property is a covariance property. Suppose that $\psi$ is
1417: a diffeomorphism from the manifold $M$ to another manifold $N$ that
1418: takes ${\cal P}$ to $\psi({\cal P})$. We denote by $\psi^*$ the
1419: natural mapping of tensors over the tangent space at ${\cal P}$ to
1420: tensors over the tangent space at $\psi({\cal P})$. We also denote by
1421: $\psi^*$ the associated natural mapping of tensor fields on $M$ to
1422: tensor fields on $N$. Then the covariance property is
1423: \be
1424: f^\alpha[\psi({\cal P}), \psi^* u^\alpha, \psi^* g_{\alpha\beta},
1425: \psi^* \Phi] = \psi^* f^\alpha[{\cal P}, u^\alpha, g_{\alpha\beta},
1426: \Phi].
1427: \ee
1428: This expresses the fact that the self-force does not depend on
1429: quantities such as a choice of coordinates.
1430: It follows from the definition (\ref{eq:selfret}) that the retarded
1431: self-force $f^\alpha_{\rm ret}$ satisfies a similar covariance relation:
1432: \be
1433: f^\alpha_{\rm ret}[\psi({\cal P}), \psi^* u^\alpha, \psi^* g_{\alpha\beta}
1434: ] = \psi^* f^\alpha_{\rm ret}[{\cal P}, u^\alpha, g_{\alpha\beta}].
1435: \label{eq:covariance}
1436: \ee
1437:
1438:
1439:
1440: \subsection{Property of generic bound geodesics in Kerr}
1441:
1442: Next we review a property of bound geodesics in Kerr upon which the
1443: proof depends. This is the fact that
1444: for generic bound geodesics there exists isometries $\psi$
1445: of the Kerr spacetime of the form
1446: \be
1447: t \to 2 t_1 - t, \ \ \ \ \phi \to 2 \phi_1 - \phi,
1448: \label{eq:isometry}
1449: \ee
1450: where $t_1$ and $\phi_1$ are constants,
1451: which come arbitrarily close to mapping the geodesic onto itself.
1452:
1453: To see this, note that if a geodesic is mapped onto itself by the
1454: mapping (\ref{eq:isometry}), then the point
1455: $[t,r(t),\theta(t),\phi(t)]$ is mapped onto
1456: $[2 t_1 - t, r(t), \theta(t),2 \phi_1 -
1457: \phi(t)]$,
1458: which must equal $[t',r(t'),\theta(t'),\phi(t')]$
1459: where $t' = 2 t_1 -t$.
1460: If we specialize to $t = t_1$ then we see that $\phi(t_1) =
1461: \phi_1$. We denote this point by ${\cal Q} =
1462: (t_1,r_1,\theta_1,\phi_1)$. Next, since the geodesic is determined
1463: by ${\cal Q}$ and by the four-velocity $u^\alpha$ at ${\cal
1464: Q}$, the geodesic is mapped onto itself by $\psi$ if and only if
1465: $\psi^* u^\alpha = - u^\alpha$ at ${\cal Q}$. This will be true if
1466: and only
1467: if the $r$ and $\theta$ components of $u^\alpha$ vanish. Thus, a
1468: geodesic will be invariant under a map of the form (\ref{eq:isometry})
1469: if and only if it contains a point ${\cal Q}$ which is simultaneously
1470: a turning point of the $r$ and $\theta$ motions.
1471:
1472:
1473: Consider now a generic bound geodesic. Such a geodesic can be
1474: characterized by the parameters $E$, $L_z$, $Q$,
1475: $\lambda_{r0}$, $\lambda_{\theta0}$, $t_0$, $\phi_0$
1476: in the notation of Sec.\ \ref{sec:geodesicparameters}.
1477: The turning points of the $r$ motion occur at
1478: $\lambda = \lambda_{r,n} \equiv \lambda_{r0} + n \Lambda_r/2$, where
1479: $n$ is an integer.
1480: Similarly the turning points of the $\theta$ motion occur at
1481: $\lambda = \lambda_{\theta,m} \equiv \lambda_{\theta0} + m
1482: \Lambda_\theta/2$, where $m$ is an
1483: integer. Since generically $\Lambda_r$ and $\Lambda_\theta$ will be
1484: incommensurate, for any $\varepsilon>0$ we can find values of the
1485: integers $m$ and $n$ so that
1486: \be
1487: |\lambda_{r,n} - \lambda_{\theta,m}| \le \varepsilon.
1488: \ee
1489: Thus we can find a point on the geodesic that is arbitrarily to close
1490: to being a turning point for both the $r$ and $\theta$ motions, and hence
1491: there is an isometry of the form (\ref{eq:isometry})
1492: that comes arbitrarily close to mapping the geodesic onto itself.
1493:
1494: Below we will simply assume that the isometry does map the geodesic
1495: onto itself; the error associated with this assumption can be made
1496: arbitrarily small in the adiabatic limit\footnote{Note however that
1497: real inspirals will consist of finite curves which are approximately
1498: locally geodesic; the corresponding error for such curves will
1499: contribute to the post-adiabatic correction to the inspiral, i.e., to
1500: the second term in Eq.\ (\protect{\ref{eq:emri_phase}}).}.
1501:
1502:
1503:
1504:
1505: \subsection{Property of the adiabatic limit}
1506:
1507: The self-acceleration $a^\alpha$ at a point ${\cal P}$ on a geodesic
1508: has three independent components, since $a^\alpha u_\alpha =0$. These
1509: three components are determined by the time derivatives of the three
1510: conserved quantities
1511: $$
1512: {dE \over dt},\ \ \
1513: {dL_z \over dt},\ \ \
1514: {dK \over dt}.
1515: $$
1516: In order to compute the evolution of the orbit in the adiabatic limit,
1517: it is sufficient to know the time averaged quantities
1518: \be
1519: \left< {dE \over dt} \right>_t,\ \ \
1520: \left< {dL_z \over dt} \right>_t,\ \ \
1521: \left< {dK \over dt} \right>_t,
1522: \label{eq:timeaverages}
1523: \ee
1524: where
1525: \be
1526: \left< {dE \over dt} \right>_t \equiv \lim_{T \to \infty} {1 \over 2
1527: T} \int_{-T}^T dt \, {dE \over dt}(t).
1528: \ee
1529: This was shown in Ref.\ \cite{Mino:2003yg}
1530: (although see Refs.\ \cite{2005PThPh.113..733M,2005Minob,2005Minoc}
1531: for caveats related to gauge issues), and is
1532: also discussed in detail using a two-timescale expansion in Ref.\
1533: \cite{Hinderer}.
1534:
1535:
1536: \subsection{Derivation}
1537:
1538: We now combine the various results discussed above to derive the half
1539: retarded minus half advanced prescription.
1540: We parameterize the geodesic as $x^\alpha = z^\alpha(t)$, and denote
1541: by
1542: \be
1543: f^{\alpha}_{\rm ret}(t) = f_{\rm ret}^\alpha[ z^\alpha(t),
1544: u^\alpha(t), g_{\alpha\beta} ]
1545: \ee
1546: the retarded self force at $z^\alpha(t)$.
1547: The key idea is to use the covariance property (\ref{eq:covariance}),
1548: with $\psi$ taken to be the isometry associated with the geodesic.
1549: If we take ${\cal P}$ to be $z^\alpha(t)$, then the
1550: left hand side of (\ref{eq:covariance}) evaluates to
1551: \begin{eqnarray}
1552: \fl
1553: f_{\rm ret}^\alpha[ z^\alpha(2 t_1 -t), - u^\alpha(2 t_1 - t),
1554: g_{\alpha\beta}] &=& f_{\rm adv}^\alpha[ z^\alpha(2 t_1 -t), u^\alpha(2 t_1 - t),
1555: g_{\alpha\beta}] \nn \\
1556: &=& f_{\rm adv}^\alpha(2 t_1 - t).
1557: \end{eqnarray}
1558: Here we have used the property (\ref{eq:flip}) and the fact that
1559: $\psi^* g_{\alpha\beta} = g_{\alpha\beta}$.
1560: This gives
1561: \be
1562: f_{\rm adv}^\alpha(2 t_1 -t ) = \psi^* f_{\rm ret}^\alpha(t).
1563: \label{eq:flip1}
1564: \ee
1565:
1566: Now let ${\cal E}$ denote one of the three conserved quantities $E$,
1567: $L_z$ or $K$. The time derivative of ${\cal E}$ depends linearly on
1568: the self force, so we can write
1569: \be
1570: {d {\cal E} \over d t}(t) = \Xi_\alpha(t) f^\alpha_{\rm ret}(t)
1571: \label{eq:dcalEdt000}
1572: \ee
1573: for some $\Xi_\alpha$.
1574: Below we will compute $\Xi_\alpha$ explicitly for the three different
1575: conserved quantities and derive the key property that
1576: \be
1577: \psi^* \Xi_\alpha(t) = - \Xi_\alpha(2 t_1 -t).
1578: \label{eq:flip2}
1579: \ee
1580: Now acting on Eq. (\ref{eq:dcalEdt000}) with $\psi^*$ and using the
1581: relations (\ref{eq:flip1}) and (\ref{eq:flip2})
1582: yields
1583: \be
1584: (\Xi_\alpha f^\alpha_{\rm ret})(t) = - (\Xi_\alpha f^\alpha_{\rm
1585: adv})(2t_1 -t).
1586: \ee
1587: Taking a time average gives
1588: \be
1589: \left< \Xi_\alpha f_{\rm ret}^\alpha \right>_t = -
1590: \left< \Xi_\alpha f_{\rm adv}^\alpha \right>_t,
1591: \ee
1592: and hence
1593: \be
1594: \left< { d {\cal E} \over dt } \right>_t =
1595: \left< \Xi_\alpha f_{\rm ret}^\alpha \right>_t =
1596: {1 \over 2} \left<
1597: \Xi_\alpha ( f^\alpha_{\rm ret} - f^\alpha_{\rm adv}) \right>_t.
1598: \ee
1599: Finally we use the explicit formula (\ref{eq:property0}) for the self force. The
1600: contributions from the locally constructed singular field $\Phi_{\rm
1601: sing}$ cancel, and we obtain
1602: \be
1603: \left< { d {\cal E} \over dt } \right>_t =
1604: {q \over 2} \left<
1605: \Xi_\alpha \nabla^\alpha ( \Phi_{\rm ret} - \Phi_{\rm adv}) \right>_t.
1606: \ee
1607: This can be written as
1608: \be
1609: \left< { d {\cal E} \over dt } \right>_t = \left<
1610: \Xi_\alpha f_{\rm rad}^\alpha \right>_t
1611: \ee
1612: where
1613: \be
1614: f_{\rm rad}^\alpha = {q \over 2} \nabla^\alpha ( \Phi_{\rm ret} -
1615: \Phi_{\rm adv} )
1616: \ee
1617: is the self-force one obtains from the half retarded minus half
1618: advanced prescription, cf.\ Eqs.\ (\ref{eq:selfacc0}) and
1619: (\ref{eq:dmudt2}) above.
1620: What we have shown is that although the quantity $f_{\rm rad}^\alpha$
1621: differs from the true self-force
1622: $f_{\rm ret}^\alpha$, the difference averages to zero when one
1623: takes a time average over the entire geodesic.
1624:
1625: It remains to derive the formula (\ref{eq:flip2}).
1626: For energy and angular momentum, we can write the conserved quantity as
1627: the inner product of a Killing vector $\xi^\alpha$ with the 4-velocity
1628: of the particle:
1629: \be
1630: {\cal E} = \xi_\alpha u^\alpha.
1631: \ee
1632: Here ${\vec \xi} = - \partial/\partial t$ for ${\cal E} = E$, and
1633: ${\vec \xi} = \partial /\partial \phi$ for ${\cal E} = L_z$.
1634: Taking a derivative with respect to proper time $\tau$ gives
1635: \be
1636: \frac{d {\cal E}}{d \tau} = u^\beta \nabla_\beta ( \xi_\alpha
1637: u^\alpha) = (\nabla_\beta \xi_\alpha) u^\beta u^\alpha + \xi_\alpha
1638: (u^\beta \nabla_\beta u^\alpha) = \xi_\alpha a^\alpha,
1639: \ee
1640: where $a^\alpha$ is the 4-acceleration. Here we have used the fact
1641: that ${\vec \xi}$ is a Killing vector so $\nabla_{(\alpha}
1642: \xi_{\beta)} =0$. Next we use the geodesic equations (\ref{eq:tdot})
1643: and (\ref{eq:Minotime}) to
1644: translate from a proper time derivative to coordinate time derivative,
1645: and we use the relation between self-force and self-acceleration given
1646: by Eqs.\ (\ref{eq:acc}) and (\ref{eq:tsf}). Using the definition
1647: (\ref{eq:dcalEdt000}) of $\Xi_\alpha$ now gives
1648: \be
1649: \Xi_\alpha = \frac{\Sigma(r)}{\mu V_t(r,\theta)} (\delta^\beta_\alpha +
1650: u_\alpha u^\beta) \xi_\beta.
1651: \ee
1652: The formula (\ref{eq:flip2}) now follows from $\psi^* u^\alpha(t) = -
1653: u^\alpha(2 t_1 -t)$ and $\psi^* \xi_\alpha = - \xi_\alpha$.
1654:
1655:
1656: Turn now to the Carter constant.
1657: Taking a time derivative of the expression (\ref{eq:Kdef}) for $K$ gives
1658: \begin{eqnarray}
1659: \frac{d K}{d \tau} &=& u^\gamma \nabla_\gamma (K_{\alpha\beta} u^\alpha u^\beta) =
1660: \nabla_{(\gamma} K_{\alpha\beta)} u^\gamma u^\alpha u^\beta + 2 K_{\alpha\beta} u^\alpha u^\gamma \nabla_\gamma u^\beta \nn \\
1661: \mbox{} &=& 2 K_{\alpha\beta} u^\alpha a^\beta.
1662: \end{eqnarray}
1663: Here we have used the Killing tensor equation $\nabla_{(\gamma} K_{\alpha\beta)} =0$.
1664: Proceeding as before now gives the expression
1665: \be
1666: \Xi_\alpha = \frac{2 \Sigma(r)}{\mu V_t(r,\theta)} (\delta^\beta_\alpha +
1667: u_\alpha u^\beta) K_{\beta\gamma} u^\gamma.
1668: \ee
1669: The formula (\ref{eq:flip2}) now follows from the relations $\psi^* u^\alpha(t) = -
1670: u^\alpha(2 t_1 -t)$ and $\psi^* K_{\alpha\beta} = K_{\alpha\beta}$.
1671:
1672:
1673:
1674: \section{Separation of variables and basis of modes}
1675: \label{sec:modes}
1676:
1677: \subsection{Separation of variables for scalar wave equation}
1678:
1679: Separation of variables for the scalar wave equation in Kerr was first
1680: carried out by Dieter Brill and others in 1972 \cite{Brill:1972xj},
1681: and subsequently generalized to higher spins by Teukolsky
1682: \cite{Teukolsky:1972my,Teukolsky:1973ha}. We substitute the ansatz
1683: \be
1684: \Phi(t,r,\theta,\phi) = R(r) \Theta(\theta) e^{i m \phi} e^{-i \omega
1685: t}
1686: \label{eq:ansatz}
1687: \ee
1688: into the homogeneous version of Eq.\ (\ref{eq:wave}), and make use of the
1689: expression (\ref{eq:waveoperator}) for the wave operator. The result
1690: is the two equations
1691: \be
1692: \frac{1}{\sin \theta} (\sin \theta \Theta_{,\theta})_{,\theta} -
1693: \frac{m^2}{\sin^2 \theta} \Theta + a^2 \omega^2 \cos^2\theta \Theta =
1694: - \lambda \Theta,
1695: \label{eq:Thetaeqn}
1696: \ee
1697: and
1698: \be
1699: ( \Delta R_{,r})_{,r} + \frac{1}{\Delta} \left( \omega^2 \varpi^4 - 4
1700: r a m \omega + m^2 a^2 \right) R = (\lambda + \omega^2 a^2) R.
1701: \label{eq:Reqn}
1702: \ee
1703: Here $\lambda$ is the separation constant
1704: We label the successive
1705: eigenvalues $\lambda$ of Eq.\ (\ref{eq:Thetaeqn}) by an integer $l$,
1706: with $l = |m|,|m|+1,|m|+2, \ldots$, and we denote these eigenvalues by
1707: $\lambda_{\omega lm}$.
1708: In the Schwarzschild limit $a=0$ we have $\lambda_{\omega lm} = l(l+1)$.
1709: Since the differential equation
1710: (\ref{eq:Thetaeqn}) does not depend on the sign of $\omega$ or the
1711: sign of $m$, we have
1712: \be
1713: \lambda_{-\omega l-m} = \lambda_{\omega lm},
1714: \label{eq:lambdaidentity}
1715: \ee
1716: a property which will be used later.
1717: We denote by $\Theta_{\omega lm}(\theta)$ the corresponding solutions to
1718: Eq.\ (\ref{eq:Thetaeqn}). The functions
1719: \be
1720: S_{\omega lm}(\theta,\phi) \equiv e^{i m \phi} \Theta_{\omega lm}(\theta)
1721: \label{eq:Somegalmdef}
1722: \ee
1723: are the spheroidal harmonics.
1724: These harmonics are orthogonal on the sphere, and we can choose them
1725: to be orthonormal like the spherical harmonics:
1726: \be
1727: \int d^2 \Omega \ S_{\omega lm}(\theta,\phi)^* S_{\omega
1728: l'm'}(\theta,\phi) = \delta_{l l'} \delta_{m m'}.
1729: \label{eq:orthonormal}
1730: \ee
1731: Following Gal'tsov \cite{Galtsov:1982} we also choose the phases of the
1732: spheroidal harmonics to satisfy
1733: \be
1734: (P S_{\omega lm})(\theta,\phi) \equiv S_{\omega lm}(\pi - \theta, \pi +
1735: \phi) = (-1)^l S_{\omega lm}(\theta,\phi)
1736: \label{eq:Zparity}
1737: \ee
1738: and
1739: \be
1740: S_{-\omega l -m}(\theta,\phi)^* = (-1)^m S_{\omega lm}(\theta,\phi).
1741: \label{eq:Zparity1}
1742: \ee
1743: Here $P$ is the parity operator.
1744:
1745:
1746: The radial equation (\ref{eq:Reqn}) can now be simplified by
1747: defining
1748: \be
1749: u(r) = \varpi R(r),
1750: \label{eq:udef}
1751: \ee
1752: and by using the tortoise coordinate
1753: $r^*$ defined by
1754: \be
1755: dr^* = (\varpi^2/\Delta) dr.
1756: \label{eq:rstardef0}
1757: \ee
1758: An explicit
1759: expression for $r^*$ is
1760: \be
1761: r^* = r + \frac{2 r_+}{ r_+ - r_-}\ln \frac{r-r_+}{ 2 }
1762: - \frac{2 r_-}{ r_+ - r_-}\ln \frac{r-r_-}{ 2 },
1763: \label{eq:rstardef}
1764: \ee
1765: where $r_\pm = 1 \pm \sqrt{1^2 - a^2}$ are the two roots of $\Delta(r)
1766: =0$. [Recall that we are using units in which the mass $M$ of the
1767: black hole is unity.]
1768: The resulting simplified radial equation is
1769: \be
1770: \frac{d^2 u}{d r^{*\,2}}(r^*) + V_{\omega lm}(r^*) u(r^*) = 0,
1771: \label{eq:diffsimple}
1772: \ee
1773: where the potential is
1774: \begin{eqnarray}
1775: \fl
1776: V_{\omega lm} = - \frac{\Delta}{\varpi^6}(3r^2 - 4 r + a^2) + \frac{3 r^2
1777: \Delta^2}{\varpi^8} + \omega^2
1778: \nonumber \\
1779: + \frac{1}{\varpi^4} \left[ m^2 a^2 -
1780: 4 r a m \omega - \Delta (\lambda_{\omega l m} + \omega^2 a^2 )
1781: \right].
1782: \label{eq:potential}
1783: \end{eqnarray}
1784: From Eq.\ (\ref{eq:lambdaidentity}) it follows that this potential satisfies
1785: \be
1786: V_{-\omega l-m}(r) = V_{\omega lm}(r),
1787: \label{eq:Videntity}
1788: \ee
1789: an identity which will be used later.
1790:
1791: \subsection{Basis of modes}
1792:
1793: In this section we follow the treatment of Chrzanowski
1794: \cite{Chrzanowski:1975wv} as modified slightly by Gal'tsov \cite{Galtsov:1982}.
1795: We follow Gal'tsov's notation except that we omit
1796: his subscripts $s$ denoting spin, since we have specialized to $s=0$.
1797:
1798:
1799: We start by analyzing the asymptotic behavior of the potential
1800: (\ref{eq:potential}) as $r^* \to -\infty$ and as $r^* \to \infty$.
1801: We consider first the limit $r^* \to -\infty$, the past and future
1802: event horizons. In this limit $\Delta \to 0$ and $r \to r_+ = 1 +
1803: \sqrt{1 - a^2}$.
1804: We find $V_{\omega lm}(r) \to V_{\omega lm}(r_+) = p_{m\omega}^2$,
1805: where
1806: \be
1807: p_{m\omega} = \omega - m \omega_+,
1808: \label{eq:kdef}
1809: \ee
1810: and
1811: \be
1812: \omega_+ = \frac{a}{2 r_+}
1813: \ee
1814: is the angular velocity of rotation of the horizon. Therefore the
1815: solutions of the radial equation near the horizons are of the form
1816: \be
1817: u(r) \propto e^{\pm i p_{m\omega} r^*}.
1818: \ee
1819: In the limit of $r^* \to \infty$ (past and future
1820: null infinity), $V_{\omega lm}(r^*) \to \omega^2$, so the solutions are of the
1821: form
1822: \be
1823: u(r) \propto e^{\pm i \omega r^*}.
1824: \ee
1825:
1826:
1827: \subsubsection{``In'', ``up'', ``out'' and ``down'' modes}
1828:
1829: We now define, following Gal'tsov, the following solution
1830: \be
1831: u_{\omega lm}^{\rm in}(r^*) = \alpha_{\omega lm}
1832: \left\{ \begin{array}{ll} \tau_{\omega lm} |p_{m\omega}|^{-1/2} e^{- i p_{m\omega} r^*},
1833: & \mbox{ $r^* \to -\infty$}\\
1834: |\omega|^{-1/2} \left[ e^{-i \omega r^*} + \sigma_{\omega lm} e^{i
1835: \omega r^*} \right], & \mbox{
1836: $r^* \to \infty$.}\\ \end{array} \right.
1837: \label{eq:uindef}
1838: \ee
1839: This equation serves to define the mode as well as the complex
1840: transmission and reflection coefficients $\tau_{\omega lm}$ and
1841: $\sigma_{\omega lm}$. The coefficient $\alpha_{\omega lm}$ is a
1842: normalization constant whose value is arbitrary; we will discuss a
1843: convenient choice of $\alpha_{\omega lm}$ later.
1844: We will often denote the set of subscripts $\omega lm$ collectively by
1845: $\Lambda$: thus $\sigma_\Lambda$, $\tau_\Lambda$,
1846: and sometimes we will omit the subscript and use simply $\sigma$,
1847: $\tau$.
1848:
1849: The ``in'' mode (\ref{eq:uindef}) is a mixture of outgoing and ingoing
1850: components at past and future null infinity, since the mode function
1851: is multiplied by $e^{-i \omega t}$. At the past and future event
1852: horizons, the mode is purely ingoing when the sign of $p_{m\omega}$ is the same
1853: as the sign of $\omega$. However, from the definition (\ref{eq:kdef})
1854: of $p_{m\omega}$ we see that the sign of $\omega p_{m\omega}$ can be negative;
1855: this occurs
1856: for superradiant
1857: modes. Thus, at the future event horizon the ``in'' modes can be
1858: either ingoing or outgoing.
1859:
1860: \begin{figure}
1861: \begin{center}
1862: \epsfig{file=modes.eps,angle=0,width=10.5cm}
1863: \caption{An illustration of the various types of modes in black hole spacetimes.
1864: Here ${\cal J}^-$ denotes past null infinity, ${\cal J}^+$ future null
1865: infinity, $E^-$ the past event horizon, and $E^+$ the future event horizon.
1866: The four panels give the behavior of the four different modes ``in'',
1867: ``out'', ``up'' or ``down'' as indicated.
1868: A zero indicates the mode vanishes at the indicated boundary. Two
1869: arrows indicates that the
1870: mode consists of a mixture of ingoing and outgoing radiation at that
1871: boundary. Two arrows with an ``OR'' means that the mode is either
1872: purely ingoing or purely outgoing at that boundary, depending on the
1873: relative sign of $p_{m\omega}$ and $\omega$.
1874: The ``in'' modes vanish on the past event horizon, and the ``up''
1875: modes vanish on past null infinity. Thus the ``in'' and ``up'' modes
1876: together form a complete basis of modes. Similarly the ``down'' and
1877: ``out'' modes together form a complete basis of modes.}
1878: \label{fig:modes}
1879: \end{center}
1880: \end{figure}
1881:
1882:
1883:
1884: The crucial feature of the ``in'' modes is that they vanish on the
1885: past event horizon. This feature will be used later in constructing
1886: the various Green's functions. A more precise statement of the result
1887: is that a solution $\Phi$ of the wave equation which is a
1888: linear combination of ``in'' modes with coefficients $c_{\omega lm}$,
1889: such that the coefficients depend smoothly on $\omega$ (a reasonable
1890: requirement), must vanish at the past event horizon. To see this, note
1891: from Eqs.\ (\ref{eq:ansatz}) and (\ref{eq:udef})
1892: that the solution can be written as
1893: \be
1894: \Phi(t,r,\theta,\phi) = \int_{-\infty}^\infty d\omega \sum_{l=0}^\infty
1895: \sum_{m=-l}^l e^{-i \omega t} c_{\omega lm} S_{\omega lm}(\theta,\phi)
1896: \frac{u_{\omega lm}^{\rm in}(r^*)}{\varpi(r^*)}.
1897: \ee
1898: We now insert the asymptotic form (\ref{eq:uindef}) of the mode
1899: function near the horizon, and we use the definition (\ref{eq:kdef})
1900: of $p_{m\omega}$. This gives
1901: \begin{eqnarray}
1902: \fl
1903: \Phi(t,r,\theta,\phi) =
1904: \frac{1}{\varpi} \int_{-\infty}^\infty d\omega \sum_{lm} e^{-i \omega (t + r^*)}
1905: c_{\omega lm} S_{\omega lm}(\theta,\phi) \alpha_{\omega lm}
1906: \tau_{\omega lm} |p_{m\omega}|^{-1/2} e^{i m \omega_+ r^*} \\
1907: \fl
1908: \ \ \ \ \ \ \ \ \ \ \ \ \ \
1909: \mbox{} \equiv \frac{1}{\varpi} \sum_{lm} G_{lm}(t+r^*;\theta,\phi)
1910: e^{i m \omega_+ r^*}.
1911: \label{eq:Gdef}
1912: \end{eqnarray}
1913: Now all of the quantities that depend on $\omega$ in the integrand are
1914: smooth functions of $\omega$. Since Fourier transforms of smooth
1915: functions go to zero at infinity,
1916: it follows that the function $G_{lm}(v;\theta,\phi)$ defined by
1917: Eq.\ (\ref{eq:Gdef}) satisfies $G_{lm} \to 0$ as $v \to -\infty$,
1918: where $v=t+r^*$. Thus,
1919: $\Phi$ will vanish as $v\to -\infty$, on the past event horizon.
1920:
1921:
1922:
1923: Another important property of the ``in'' modes is the identity
1924: \be
1925: u^{\rm in}_{-\omega l-m}(r^*) = u^{\rm in}_{\omega lm}(r^*)^*.
1926: \label{eq:uinidentity}
1927: \ee
1928: To derive this, note that the potential $V_{\omega lm}$ is real and
1929: also satisfies the identity (\ref{eq:Videntity}). It follows that the
1930: function $u^{\rm in}_{-\omega l-m}(r^*)^*$ satisfies the same differential
1931: equation (\ref{eq:diffsimple}) as $u^{\rm in}_{\omega lm}(r)$, and has
1932: the same asymptotic behavior as $r^* \to \infty$ if we choose the
1933: normalization constants to satisfy
1934: \be
1935: \alpha_{-\omega l-m} = \alpha_{\omega lm}^*.
1936: \label{eq:alphacondt}
1937: \ee
1938: The result (\ref{eq:uinidentity}) now follows. In particular the
1939: identity (\ref{eq:uinidentity})
1940: implies that
1941: \be
1942: \sigma_{-\omega l-m} = \sigma_{\omega lm}^*
1943: \ee
1944: and
1945: \be
1946: \tau_{-\omega l-m} = \tau_{\omega lm}^*.
1947: \label{eq:tauidentity}
1948: \ee
1949:
1950:
1951:
1952:
1953: We next define the ``up'' modes
1954: \be
1955: \fl
1956: u_{\omega lm}^{\rm up}(r^*) = \beta_{\omega lm}
1957: \left\{ \begin{array}{ll} |p_{m\omega}|^{-1/2} \frac{\omega p_{m\omega}}{|\omega p_{m\omega}|}
1958: \left[ \mu_{\omega lm} e^{ i p_{m\omega} r^*} + \nu_{\omega lm} e^{-i p_{m\omega} r^*} \right],
1959: & \mbox{ $r^* \to -\infty$}\\
1960: |\omega|^{-1/2} e^{i \omega r^*}, & \mbox{
1961: $r^* \to \infty$.}\\ \end{array} \right.
1962: \label{eq:uupdef}
1963: \ee
1964: As before this equation defines both the mode as well as the complex
1965: coefficients $\mu_{\omega lm}$ and
1966: $\nu_{\omega lm}$. The coefficient $\beta_{\omega lm}$ is a
1967: normalization constant whose value is arbitrary; we will discuss a
1968: convenient choice of $\beta_{\omega lm}$ later.
1969: The ``up'' modes are a mixture of ingoing and outgoing components at the
1970: past and future event horizons. At future null infinity, the mode is
1971: purely outgoing. By using an argument similar to that given above,
1972: one can show that these modes vanish at past null infinity.
1973: Also we can argue as before that
1974: \be
1975: u_{-\omega l-m}^{\rm up}(r^*) = u_{\omega lm}^{\rm up}(r^*)^*,
1976: \label{eq:uupidentity}
1977: \ee
1978: \be
1979: \mu_{-\omega l-m} = \mu_{\omega lm}^*,
1980: \ee
1981: and
1982: \be
1983: \nu_{-\omega l-m} = \nu_{\omega lm}^*,
1984: \ee
1985: as long as we choose the normalization constants to satisfy
1986: \be
1987: \beta_{-\omega l-m} = \beta_{\omega lm}^*.
1988: \label{eq:betacondt}
1989: \ee
1990:
1991:
1992:
1993:
1994:
1995: Since the ``in'' modes vanish at the past event horizon, and the ``up'' modes
1996: vanish at past null infinity, the two sets of modes are orthogonal to
1997: one another and together form a complete basis of modes.
1998: %They are
1999: %related to the ``H'' and ``$\infty$'' modes of Hughes \cite{Hughes:1999bq} by
2000: %\be
2001: %R^H_{\omega lm}(r) \propto \frac{ u^{\rm in}_{\omega lm}}{\varpi}
2002: %\ee
2003: %and
2004: %\be
2005: %R^\infty_{\omega lm}(r) \propto \frac{ u^{\rm up}_{\omega lm}}{\varpi}.
2006: %\ee
2007:
2008: Next we note that $(u^{\rm in}_{\omega lm})^*$ is a solution of the
2009: differential equation (\ref{eq:diffsimple}) since the potential
2010: (\ref{eq:potential}) is real. We denote this quantity by
2011: \be
2012: u^{\rm out}_{\omega lm} = (u^{\rm in}_{\omega lm})^*.
2013: \label{eq:uoutdef}
2014: \ee
2015: Similarly we define
2016: \be
2017: u^{\rm down}_{\omega lm} = (u^{\rm up}_{\omega lm})^*.
2018: \label{eq:udowndef}
2019: \ee
2020: These ``down'' and ``out'' modes together form a basis, as
2021: the ``out'' modes vanish on the future event horizon and the ``down''
2022: modes vanish at future null infinity.
2023: The properties of all of these modes are summarized in Fig. \ref{fig:modes}.
2024:
2025: \subsubsection{Relations between scattering and transmission
2026: coefficients}
2027:
2028: The four sets of constants $\sigma_{\omega lm}$, $\tau_{\omega lm}$,
2029: $\mu_{\omega lm}$ and $\nu_{\omega lm}$ are not all independent.
2030: One can derive relations between these coefficients by using the fact
2031: that the Wronskian
2032: \be
2033: W(u_1,u_2) = u_1 \frac{d u_2}{d r^*} - \frac{d u_1}{d r^*} u_2.
2034: \label{eq:wronskiandef}
2035: \ee
2036: is conserved for any two solutions $u_1$
2037: and $u_2$ of the homogeneous equation (\ref{eq:diffsimple}).
2038: By evaluating $W(u^{\rm up}_{\omega lm},u^{\rm in}_{\omega lm})$
2039: at $r^* = -\infty$ and at $r^* = \infty$ using
2040: the asymptotic relations (\ref{eq:uindef}) and (\ref{eq:uupdef}) and
2041: equating the results, we obtain
2042: \be
2043: \tau_{\omega lm} \mu_{\omega lm}=1.
2044: \label{eq:widentity1}
2045: \ee
2046: A similar calculation with the modes $u^{\rm up}_{\omega lm}$ and
2047: $u^{\rm out}_{\omega lm}$ yields
2048: \be
2049: \tau_{\omega lm}^{*} \nu_{\omega lm} = - \sigma_{\omega lm}^*.
2050: \label{eq:widentity2}
2051: \ee
2052: Finally, using the modes $u^{\rm in}_{\omega lm}$ and
2053: $u^{\rm out}_{\omega lm}$ yields
2054: \be
2055: | \sigma_{\omega lm}|^2 + \frac{\omega p_{m\omega}}{|\omega
2056: p_{m\omega}|} | \tau_{\omega lm}|^2 =1.
2057: \label{eq:widentity3}
2058: \ee
2059:
2060:
2061:
2062:
2063: \subsubsection{Relations between mode functions}
2064:
2065:
2066: The ``down'' and ``out'' modes can be expressed as linear combinations
2067: of the the ``in'' and ``up'' modes, since the latter form a basis of
2068: modes. Using the asymptotic forms (\ref{eq:uindef}) and (\ref{eq:uupdef}) of the modes
2069: at $r^* \to \infty$ together with the definition (\ref{eq:udowndef})
2070: allows us to identity the coefficients for $u^{\rm down}_{\omega lm}$, giving
2071: \be
2072: u^{\rm down}_{\omega lm}(r) = \frac{\beta_{\omega lm}^*}{\alpha_{\omega lm}}
2073: u^{\rm in}_{\omega lm}(r) - \sigma_{\omega lm}
2074: \frac{\beta_{\omega lm}^*}{\beta_{\omega lm}} u^{\rm up}_{\omega
2075: lm}(r).
2076: \label{eq:udownformula}
2077: \ee
2078: A similar computation using the definition (\ref{eq:uoutdef}) and
2079: using the asymptotic forms at $r^* \to -\infty$ gives
2080: \be
2081: u^{\rm out}_{\omega lm}(r) = \frac{\omega p_{m\omega}}{|\omega p_{m\omega}|}
2082: \frac{\tau_{\omega lm}^*}{\mu_{\omega lm}} \frac{\alpha^*_{\omega
2083: lm}}{\beta_{\omega lm}} u^{\rm up}_{\omega lm}(r)
2084: - \frac{\tau_{\omega lm}^* \nu_{\omega lm}}{\mu_{\omega lm}
2085: \tau_{\omega lm}} \frac{\alpha_{\omega lm}^*}{\alpha_{\omega lm}}
2086: u^{\rm in}_{\omega lm}(r).
2087: \ee
2088: This can be simplified using the relations (\ref{eq:widentity1}) and
2089: (\ref{eq:widentity2}) to give
2090: \be
2091: u^{\rm out}_{\omega lm}(r) = \frac{\omega p_{m\omega}}{|\omega p_{m\omega}|}
2092: \frac{\alpha^*_{\omega lm}}{\beta_{\omega lm}} | \tau_{\omega lm}|^2
2093: u^{\rm up}_{\omega lm}(r)
2094: + \frac{\alpha_{\omega lm}^*}{\alpha_{\omega lm}} \sigma_{\omega lm}^*
2095: u^{\rm in}_{\omega lm}(r).
2096: \label{eq:uoutformula}
2097: \ee
2098:
2099:
2100:
2101: \subsubsection{Complete mode functions}
2102:
2103: Finally, following Gal'tsov \cite{Galtsov:1982}, we define the complete
2104: mode functions
2105: \be
2106: \pi^{\rm in}_{\omega lm}(t,r,\theta,\phi) = \frac{2}{\varpi} e^{-i \omega t}
2107: S_{\omega lm}(\theta,\phi) u^{\rm in}_{\omega lm}(r^*),
2108: \label{eq:piindef}
2109: \ee
2110: \be
2111: \pi^{\rm up}_{\omega lm}(t,r,\theta,\phi) = \frac{2}{\varpi} e^{-i \omega t}
2112: S_{\omega lm}(\theta,\phi) u^{\rm up}_{\omega lm}(r^*),
2113: \label{eq:piupdef}
2114: \ee
2115: \be
2116: \pi^{\rm out}_{\omega lm}(t,r,\theta,\phi) = \frac{2}{\varpi} e^{-i \omega t}
2117: S_{\omega lm}(\theta,\phi) u^{\rm out}_{\omega lm}(r^*),
2118: \label{eq:pioutdef}
2119: \ee
2120: and
2121: \be
2122: \pi^{\rm down}_{\omega lm}(t,r,\theta,\phi) = \frac{2}{\varpi} e^{-i \omega t}
2123: S_{\omega lm}(\theta,\phi) u^{\rm down}_{\omega lm}(r^*).
2124: \label{eq:pidowndef}
2125: \ee
2126: Gal'tsov actually includes a factor of $1 + p P$ in these definitions,
2127: where $P$ is the parity operator defined by
2128: \be
2129: (Pf)(t,r,\theta,\phi) = f(t,r,\pi-\theta,\pi+\phi)
2130: \ee
2131: for any function $f$ and $p = \pm 1$ is an additional mode index.
2132: However when acting on
2133: the scalar modes considered in this paper, we have from Eq.\
2134: (\ref{eq:Zparity}) that
2135: \be
2136: 1 + pP = 1 + p (-1)^l = 2 \delta_{p,(-1)^l}.
2137: \ee
2138: Thus the $p$ index is redundant in the scalar case; we have simply
2139: dropped it and replaced the factor of $1+pP$ by $2$.
2140:
2141:
2142: \section{Construction of the retarded Green's function}
2143: \label{sec:retarded}
2144:
2145: \subsection{Formulae for retarded Green's function}
2146: \label{sec:retardedformula}
2147:
2148: The retarded Green's function $G_{\rm ret}(x,x')$ is defined such that if $\Phi$
2149: obeys the scalar wave equation with source ${\cal T}$
2150: \be
2151: \Box \Phi = {\cal T},
2152: \ee
2153: then the retarded solution is
2154: \be
2155: \Phi_{\rm ret}(x) = \int d^4 x' \sqrt{-g(x')} G_{\rm ret}(x,x') {\cal T}(x').
2156: \label{eq:Gretdef}
2157: \ee
2158: The expression for the retarded Green's function in terms of the modes
2159: defined in the last section is\footnote{This expression agrees with
2160: Eq.\ (2.17) of Gal'tsov \protect{\cite{Galtsov:1982}},
2161: if we assume that the normalization constants obey
2162: $
2163: \alpha_{\omega lm} \beta_{\omega lm} = - 1/(16 \pi).
2164: $
2165: However Eq.\ (2.25) of Gal'tsov
2166: says $\alpha_{\omega lm} \beta_{\omega lm} = 1/4$; thus we disagree
2167: with Gal'tsov by a factor of $-4 \pi$. For this reason we do not
2168: assume any values for the normalization constants $\alpha_{\omega lm}$
2169: and $\beta_{\omega lm}$ in our computations, but leave them
2170: unevaluated.}
2171: \begin{eqnarray}
2172: \fl
2173: G_{\rm ret}(x,x') = \frac{1}{16 \pi i} \int_{-\infty}^\infty d\omega
2174: \sum_{l=0}^\infty \sum_{m=-l}^l \frac{\omega}{|\omega|}
2175: \frac{1}{\alpha_{\omega lm} \beta_{\omega lm}} \nonumber \\
2176: \lo \times
2177: \left[
2178: \pi^{\rm up}_{\omega lm}(x) \pi^{\rm out}_{\omega lm}(x')^* \theta(r-r') +
2179: \pi^{\rm in}_{\omega lm}(x) \pi^{\rm down}_{\omega lm}(x')^* \theta(r'-r)
2180: \right].
2181: \label{eq:Gretformula0}
2182: \end{eqnarray}
2183: Here $\theta(x)$ is the step function, defined to be $+1$ for $x \ge
2184: 0$ and $0$ otherwise.
2185: Note that the expression (\ref{eq:Gretformula0}) is
2186: independent of the values chosen for the normalization
2187: constants $\alpha_{\omega lm}$ and $\beta_{\omega lm}$, since the
2188: factor of $1/\alpha$ cancels a factor of $\alpha$
2189: present in the definition (\ref{eq:uindef}) of the ``in'' modes, and
2190: similarly for $\beta$ and the ``up'' modes.
2191:
2192:
2193:
2194: The expression (\ref{eq:Gretformula0})
2195: can be expanded into a more
2196: explicit form by using the definitions (\ref{eq:piindef}) --
2197: (\ref{eq:pidowndef}) of the complete mode functions $\pi_{\omega
2198: lm}(t,r,\theta,\phi)$ in terms of the radial mode functions
2199: $u_{\omega lm}(r)$, together with the definitions (\ref{eq:uoutdef})
2200: and (\ref{eq:udowndef}) of the ``out'' and ``down'' modes.
2201: This gives
2202: \begin{eqnarray}
2203: \fl
2204: G_{\rm ret}(t,r,\theta,\phi;t',r',\theta',\phi') = \frac{1}{4 \pi i}
2205: \int_{-\infty}^\infty d\omega e^{-i \omega (t - t')}
2206: \sum_{lm} S_{\omega lm}(\theta,\phi) S_{\omega
2207: lm}(\theta',\phi')^*
2208: \frac{\omega}{|\omega|}
2209: \nn \\
2210: \fl \ \ \ \ \ \ \times
2211: \frac{1}{\varpi \varpi' \alpha_{\omega lm} \beta_{\omega lm}}
2212: \left[ u^{\rm up}_{\omega lm}(r) u^{\rm in}_{\omega lm}(r') \theta(r-r') +
2213: u^{\rm in}_{\omega lm}(r) u^{\rm up}_{\omega lm}(r') \theta(r'-r)
2214: \right].
2215: \label{eq:Gretformula1}
2216: \end{eqnarray}
2217: In this section we review the standard derivation of the formula
2218: (\ref{eq:Gretformula1}).
2219:
2220: \subsection{Derivation}
2221:
2222: The key idea behind the derivation is the following. Suppose that the
2223: source ${\cal T}(x)$ is non-zero only in the finite range of values of $r$
2224: \be
2225: r_{\rm min} \le r \le r_{\rm max}.
2226: \ee
2227: Then, the retarded solution $\Phi_{\rm ret}(x)$ will be a solution of the
2228: homogeneous equation in the regions $r < r_{\rm min}$ and $r > r_{\rm
2229: max}$. Now, the retarded solution is determined uniquely by the
2230: condition that it vanish on the past event horizon and on past null
2231: infinity. This property will be guaranteed if:
2232: \begin{enumerate}
2233: \item When we expand $\Phi_{\rm ret}$ in the region $r < r_{\rm min}$ on the
2234: basis of solutions $\pi^{\rm in}_{\omega lm}(x)$ and $\pi^{\rm
2235: up}_{\omega lm}$ of the homogeneous equation, only the ``in'' modes
2236: contribute. Then, since the ``in'' modes vanish on the past event
2237: horizon, $\Phi_{\rm ret}$ must also vanish on the past event horizon.
2238:
2239: \item When we expand $\Phi_{\rm ret}$ in the region $r > r_{\rm max}$ on the
2240: basis of solutions $\pi^{\rm in}_{\omega lm}(x)$ and $\pi^{\rm
2241: up}_{\omega lm}$, only the ``up'' modes
2242: contribute. Then, since the ``up'' modes vanish on past null
2243: infinity, $\Phi_{\rm ret}$ must also vanish on past null infinity.
2244:
2245: \end{enumerate}
2246:
2247:
2248: We start by defining the Fourier transformed quantities
2249: \be
2250: {\tilde {\cal T}}(\omega,r,\theta,\phi) = \int_{-\infty}^\infty dt e^{i
2251: \omega t} {\cal T}(t,r,\theta,\phi)
2252: \ee
2253: and
2254: \be
2255: {\tilde \Phi}(\omega,r,\theta,\phi) = \int_{-\infty}^\infty dt e^{i
2256: \omega t} \Phi(t,r,\theta,\phi).
2257: \ee
2258: [For convenience we drop the subscript ``ret'' on $\Phi$ in the
2259: remainder of this section.]
2260: We make the following ansatz for the Green's function:
2261: \be
2262: \fl
2263: G_{\rm ret}(t,r,\theta,\phi;t',r',\theta',\phi') =
2264: \int_{-\infty}^\infty \frac{d \omega}{2 \pi} e^{-i \omega (t - t')}
2265: {\tilde G}_{\rm ret}(r,\theta,\phi;r',\theta',\phi';\omega).
2266: \label{eq:ansatz0}
2267: \ee
2268: Inserting these definitions into the defining relation
2269: (\ref{eq:Gretdef}) and using the formula (\ref{eq:detg}) for $\sqrt{-g}$
2270: gives
2271: \be
2272: \fl
2273: {\tilde \Phi}(\omega,r,\theta,\phi) = \int_0^\infty dr' \int d^2
2274: \Omega' \Sigma(r',\theta') \, G_{\rm
2275: ret}(r,\theta,\phi;r',\theta',\phi';\omega) {\tilde
2276: {\cal T}}(\omega,r',\theta',\phi').
2277: \label{eq:step1}
2278: \ee
2279:
2280:
2281: Next, we decompose the quantities ${\tilde \Phi}$ and $\Sigma {\tilde
2282: {\cal T}}$ on the basis of spheroidal harmonics:
2283: \be
2284: {\tilde \Phi}(\omega,r,\theta,\phi) = \sum_{lm} S_{\omega
2285: lm}(\theta,\phi) R_{\omega lm}(r)
2286: \label{eq:Philmdef}
2287: \ee
2288: and
2289: \be
2290: \Sigma(r,\theta) {\tilde {\cal T}}(\omega,r,\theta,\phi) = r^2 \sum_{lm} S_{\omega
2291: lm}(\theta,\phi) {\tilde {\cal T}}_{\omega lm}(r)
2292: \label{eq:Tlmdef}
2293: \ee
2294: We include the factor of $\Sigma$ on the left hand side of Eq.\
2295: (\ref{eq:Tlmdef}) because of the appearance of a factor of $\Sigma$ in
2296: the integrand of Eq.\ (\ref{eq:step1}). The factor of $r^2$ on the
2297: right hand side of Eq.\ (\ref{eq:Tlmdef}) is so that the coefficients
2298: ${\tilde {\cal T}}_{\omega lm}$ reduce to the conventional spherical harmonic
2299: coefficients in the Schwarzschild limit $a =0$.
2300: From the orthogonality relation (\ref{eq:orthonormal}) the inverse
2301: transformations are
2302: \be
2303: R_{\omega lm}(r) = \int d^2 \Omega \, S_{\omega
2304: lm}(\theta,\phi)^* \, {\tilde \Phi}(\omega,r,\theta,\phi)
2305: \ee
2306: and
2307: \be
2308: r^2 {\tilde {\cal T}}_{\omega lm}(r) = \int d^2 \Omega \, S_{\omega
2309: lm}(\theta,\phi)^* \, \Sigma(r,\theta) {\tilde {\cal T}}(\omega,r,\theta,\phi).
2310: \label{eq:inverse2}
2311: \ee
2312:
2313:
2314: Next, we insert the decompositions (\ref{eq:Philmdef}) and
2315: (\ref{eq:Tlmdef}) of ${\tilde \Phi}$ and ${\tilde {\cal T}}$ into the
2316: Fourier transform of the differential equation (\ref{eq:wave}) using
2317: Eq.\ (\ref{eq:waveoperator}). This gives
2318: \be
2319: \frac{d^2 u_{\omega lm}}{d r^{*\,2}} + V_{\omega lm}(r^*) u_{\omega lm}(r^*) =
2320: s_{\omega lm},
2321: \label{eq:ulmeqn}
2322: \ee
2323: where $u_{\omega lm}(r) = \varpi(r) R_{\omega lm}(r)$, the potential
2324: $V_{\omega lm}$ is given by Eq.\ (\ref{eq:potential}), and
2325: the source term is
2326: \be
2327: s_{\omega lm} = \frac{r^2 \Delta}{\varpi^3} {\tilde {\cal T}}_{\omega lm}.
2328: \label{eq:slmdef}
2329: \ee
2330: We denote by $G_{\omega lm}(r^*,r^{*\,\prime})$ the relevant
2331: Green's function for the differential equation
2332: (\ref{eq:ulmeqn}):
2333: \be
2334: u_{\omega lm}(r^*) = \int_{-\infty}^\infty dr^{*\,\prime} G_{\omega
2335: lm}(r^*,r^{*\,\prime}) s_{\omega lm}(r^{*\,\prime}).
2336: \label{eq:Gomegalmdef}
2337: \ee
2338: We will derive an explicit formula for $G_{\omega
2339: lm}(r^*,r^{*\,\prime})$ shortly.
2340: First, we note that the Fourier-transformed retarded Green's function
2341: ${\tilde G}_{\rm ret}(r,\theta,\phi;r',\theta',\phi';\omega)$ can be
2342: expressed in terms of $G_{\omega lm}$ via the formula
2343: \be
2344: \fl
2345: {\tilde G}_{\rm ret}(r,\theta,\phi;r',\theta',\phi';\omega) =
2346: \sum_{lm} S_{\omega lm}(\theta,\phi) S_{\omega lm}(\theta',\phi')^* \,
2347: \frac{G_{\omega lm}(r^*,r^{*\,\prime})}{\varpi \varpi'}.
2348: \label{eq:ansatz1}
2349: \ee
2350: To see this, insert the ansatz (\ref{eq:ansatz1}) into the
2351: relation (\ref{eq:step1}) and simplify using the definition
2352: (\ref{eq:inverse2}) of ${\tilde {\cal T}}_{\omega lm}$. This gives
2353: \be
2354: \fl
2355: {\tilde \Phi}(\omega,r,\theta,\phi) = \sum_{lm} \int_0^\infty dr'
2356: S_{\omega lm}(\theta,\phi) \frac{r^{'\,2}}{\varpi \varpi'} G_{\omega
2357: lm}(r^*,r^{*\,\prime}) {\tilde {\cal T}}_{\omega lm}(r').
2358: \ee
2359: Comparing this with the definition (\ref{eq:Philmdef}) of $R_{\omega
2360: lm}$ and simplifying using the relations (\ref{eq:udef}),
2361: (\ref{eq:rstardef0}) and (\ref{eq:slmdef}) finally yields the formula
2362: (\ref{eq:Gomegalmdef}). Hence the ansatz (\ref{eq:ansatz1}) is correct.
2363:
2364:
2365: Finally we turn to the derivation of the formula for the
2366: Green's function $G_{\omega lm}$ for the differential equation
2367: (\ref{eq:ulmeqn}). From the discussion at the start of this section,
2368: the relevant boundary conditions to impose are that
2369: \be
2370: G_{\omega lm}(r^*,r^{*\,\prime}) \propto u^{\rm in}_{\omega lm}(r^*), \
2371: \ \ r^* \to -\infty
2372: \label{eq:b1}
2373: \ee
2374: and
2375: \be
2376: G_{\omega lm}(r^*,r^{*\,\prime}) \propto u^{\rm up}_{\omega lm}(r^*), \
2377: \ \ r^* \to \infty.
2378: \label{eq:b2}
2379: \ee
2380: Consider now the expression
2381: \begin{eqnarray}
2382: \fl
2383: G_{\omega lm}(r^*,r^{*\,\prime}) = \frac{1}{W(u^{\rm in}_{\omega lm},
2384: u^{\rm up}_{\omega lm})} \bigg[
2385: u^{\rm up}_{\omega lm}(r) u^{\rm in}_{\omega lm}(r^\prime) \theta(r-r')
2386: \nonumber \\
2387: \lo +
2388: u^{\rm in}_{\omega lm}(r) u^{\rm up}_{\omega lm}(r^\prime)
2389: \theta(r-r') \bigg],
2390: \label{eq:ansatz2}
2391: \end{eqnarray}
2392: where $W$ is the conserved Wronskian (\ref{eq:wronskiandef}).
2393: This expression satisfies the boundary conditions (\ref{eq:b1}) and
2394: (\ref{eq:b2}). Also one can directly verify that it satisfies the
2395: differential equation (\ref{eq:ulmeqn}) with the source replaced by
2396: $\delta(r^* - r^{*\,\prime})$, using the fact that the ``in'' and ``up''
2397: modes satisfy the homogeneous version of the differential equation.
2398: This establishes the formula (\ref{eq:ansatz2}).
2399:
2400: Next, we compute the Wronskian
2401: $W(u^{\rm in}_{\omega lm}, u^{\rm up}_{\omega lm})$ using the
2402: asymptotic expressions (\ref{eq:uindef}) and (\ref{eq:uupdef}) for the
2403: mode functions for $r^* \to \infty$. This gives
2404: \be
2405: W(u^{\rm in}_{\omega lm}, u^{\rm up}_{\omega lm}) = 2 i \alpha_{\omega
2406: lm} \beta_{\omega lm} \frac{\omega}{|\omega|}.
2407: \ee
2408: Inserting this into Eq.\ (\ref{eq:ansatz2}) and then into Eqs.\
2409: (\ref{eq:ansatz1}) and (\ref{eq:ansatz0}) finally yields the formula
2410: (\ref{eq:Gretformula1}).
2411:
2412:
2413: \section{Construction of the radiative Green's function}
2414: \label{sec:radiative}
2415:
2416: \subsection{Formulae for radiative Green's function}
2417:
2418:
2419: Using the retarded Green's function $G_{\rm ret}(x,x')$ discussed in the last section
2420: we can construct the retarded solution $\Phi_{\rm ret}(x)$ of the wave
2421: equation (\ref{eq:wave}). Similarly, we can construct the advanced solution
2422: $\Phi_{\rm adv}(x)$ using the advanced Green's function $G_{\rm
2423: adv}(x,x')$; this is the unique solution which vanishes on the
2424: future event horizon and at future null infinity.
2425: One half the retarded solution minus one half the advanced solution
2426: gives the radiative solution:
2427: \be
2428: \Phi_{\rm rad}(x) = \frac{1}{2} \left[ \Phi_{\rm ret}(x) - \Phi_{\rm
2429: adv}(x) \right].
2430: \ee
2431: Clearly the radiative solution is given in terms of a radiative Green's
2432: function
2433: \be
2434: \Phi_{\rm rad}(x) = \int d^4 x' \sqrt{-g(x')} G_{\rm rad}(x,x') {\cal T}(x'),
2435: \ee
2436: where
2437: \be
2438: G_{\rm rad}(x,x') = \frac{1}{2} \left[ G_{\rm ret}(x,x') - G_{\rm
2439: adv}(x,x') \right].
2440: \label{eq:Graddef}
2441: \ee
2442: The expression for the radiative Green's function in terms of the modes
2443: defined in Sec.\ \ref{sec:modes} is
2444: \begin{eqnarray}
2445: \fl
2446: G_{\rm rad}(x,x') = \frac{1}{32 \pi i} \int_{-\infty}^\infty d\omega
2447: \sum_{l=0}^\infty \sum_{m=-l}^l \frac{\omega}{|\omega|}
2448: \bigg[
2449: \frac{1}{|\alpha_{\omega lm}|^2}
2450: \pi^{\rm out}_{\omega lm}(x) \pi^{\rm out}_{\omega
2451: lm}(x')^* \nonumber \\
2452: \lo
2453: +\frac{\omega p_{m\omega}}{|\omega p_{m\omega}|}
2454: \frac{|\tau_{\omega lm}|^2}{|\beta_{\omega lm}|^2}
2455: \pi^{\rm down}_{\omega lm}(x) \pi^{\rm down}_{\omega lm}(x')^*
2456: \bigg].
2457: \label{eq:Gradformula0}
2458: \end{eqnarray}
2459: Note that this expression
2460: actually independent of the values chosen for the normalization
2461: constants $\alpha_{\omega lm}$ and $\beta_{\omega lm}$, since the
2462: factor of $1/|\alpha|^2$ cancels factors of $\alpha$
2463: present in the definition (\ref{eq:uindef}) and (\ref{eq:uoutdef}) of
2464: the ``out'' modes, and similarly for $\beta$ and the ``down'' modes.
2465:
2466:
2467: The expression (\ref{eq:Gradformula0})
2468: can be expanded into a more
2469: explicit form by using the definitions (\ref{eq:piindef}) --
2470: (\ref{eq:pidowndef}) of the complete mode functions $\pi_{\omega
2471: lm}(t,r,\theta,\phi)$ in terms of the radial mode functions
2472: $u_{\omega lm}(r)$.
2473: This gives
2474: \begin{eqnarray}
2475: \fl
2476: G_{\rm rad}(t,r,\theta,\phi;t',r',\theta',\phi') = \frac{1}{8 \pi i}
2477: \int_{-\infty}^\infty d\omega e^{-i \omega (t - t')}
2478: \sum_{lm} S_{\omega lm}(\theta,\phi) S_{\omega
2479: lm}(\theta',\phi')^*
2480: \frac{\omega}{|\omega|}
2481: \frac{1}{\varpi \varpi'} \nn \\
2482: \fl \ \ \ \ \times
2483: \left[
2484: \frac{1}{|\alpha_{\omega lm}|^2}
2485: u^{\rm out}_{\omega lm}(r) u^{\rm out}_{\omega
2486: lm}(r')^* +\frac{\omega p_{m\omega}}{|\omega p_{m\omega}|}
2487: \frac{|\tau_{\omega lm}|^2}{|\beta_{\omega lm}|^2}
2488: u^{\rm down}_{\omega lm}(r) u^{\rm down}_{\omega lm}(r')^*
2489: \right].
2490: \label{eq:Gradformula1}
2491: \end{eqnarray}
2492: In this section we review the derivation of the formula
2493: (\ref{eq:Gradformula1}) given in Gal'tsov \cite{Galtsov:1982}.
2494:
2495:
2496: \subsection{Derivation}
2497:
2498: We start by deriving the identity
2499: \be
2500: G_{\rm adv}(x,x') = G_{\rm ret}(x',x).
2501: \label{eq:identitya}
2502: \ee
2503: For any two functions $\Phi(x)$ and $\Psi(x)$ we have the identity
2504: \be
2505: \nabla^a ( \Phi \nabla_a \Psi - \Psi \nabla_a \Phi) = \Phi \Box \Psi - \Psi \Box \Phi.
2506: \ee
2507: Integrating this identity over a spacetime region $V$ and using Stokes theorem gives
2508: \be
2509: \int_{\partial V} (\Phi \nabla^a \Psi - \Psi \nabla^a \Phi) d^3 \Sigma_a = \int_V d^4 x \sqrt{-g} ( \Phi \Box \Psi - \Psi \Box \Phi).
2510: \label{eq:identity0}
2511: \ee
2512: If we choose $V$ to be the region of spacetime outside the black hole,
2513: then the boundary $\partial V$ of $V$ consists of the past and future
2514: event horizons, and also past and future null infinity.
2515:
2516:
2517: We now fix two points $x_1$ and $x_2$ in $V$, and we choose
2518: \be
2519: \Phi(x) = G_{\rm ret}(x,x_1)
2520: \ee
2521: and
2522: \be
2523: \Psi(x) = G_{\rm adv}(x,x_2).
2524: \ee
2525: Then $\Box \Phi = \delta^{(4)}(x,x_1)$ and $\Box \Psi =
2526: \delta^{(4)}(x,x_2)$, where $\delta^{(4)}$ is the covariant delta
2527: function (\ref{eq:deltafndef}). The right hand side of Eq.\
2528: (\ref{eq:identity0}) therefore evaluates to
2529: \be
2530: \Phi(x_2) - \Psi(x_1) = G_{\rm ret}(x_2,x_1) - G_{\rm adv}(x_1,x_2).
2531: \label{eq:rhsid}
2532: \ee
2533: Consider now the left hand side of Eq.\ (\ref{eq:identity0}). Since
2534: $\Phi$ vanishes on the past event horizon and at past null infinity,
2535: and $\Psi$ vanishes on the future event horizon and at future null
2536: infinity, there is no portion of $\partial V$ where both $\Phi$ and
2537: $\Psi$ are nonzero. Therefore the left hand side vanishes, so the
2538: expression (\ref{eq:rhsid}) must vanish, proving the identity
2539: (\ref{eq:identitya}).
2540:
2541:
2542: We now insert the identity (\ref{eq:identitya}) into the definition
2543: (\ref{eq:Graddef}) of $G_{\rm rad}$, and use the explicit expansion
2544: (\ref{eq:Gretformula1}) of $G_{\rm ret}$.
2545: We also specialize to
2546: \be
2547: \alpha_{\omega lm} = \beta_{\omega lm} =1;
2548: \label{eq:normalizationchoice}
2549: \ee
2550: this causes no loss in generality since the final result
2551: (\ref{eq:Gradformula1}) is
2552: independent of the values of $\alpha_{\omega lm}$ and $\beta_{\omega
2553: lm}$. This gives
2554: \begin{eqnarray}
2555: \fl
2556: G_{\rm rad}(x,x') &=&
2557: \frac{1}{8 \pi i}
2558: \int_{-\infty}^\infty d\omega e^{-i \omega (t - t')}
2559: \sum_{lm} S_{\omega lm}(\theta,\phi) S_{\omega
2560: lm}(\theta',\phi')^*
2561: \frac{\omega}{|\omega|}
2562: \frac{1}{\varpi \varpi'} \nn \\
2563: \mbox{} && \times
2564: \left[ u^{\rm up}_{\omega lm}(r) u^{\rm in}_{\omega lm}(r') \theta(r-r') +
2565: u^{\rm in}_{\omega lm}(r) u^{\rm up}_{\omega lm}(r') \theta(r'-r)
2566: \right] \nn \\
2567: \mbox{} && - \frac{1}{8 \pi i}
2568: \int_{-\infty}^\infty d\omega e^{i \omega (t - t')}
2569: \sum_{lm} S_{\omega lm}(\theta',\phi') S_{\omega
2570: lm}(\theta,\phi)^*
2571: \frac{\omega}{|\omega|}
2572: \frac{1}{\varpi \varpi'} \nn \\
2573: \mbox{} && \times
2574: \left[ u^{\rm up}_{\omega lm}(r') u^{\rm in}_{\omega lm}(r) \theta(r'-r) +
2575: u^{\rm in}_{\omega lm}(r') u^{\rm up}_{\omega lm}(r) \theta(r-r')
2576: \right],
2577: \label{eq:Gradformula3}
2578: \end{eqnarray}
2579: where $x =(t,r,\theta,\phi)$ and $x'=(t',r',\theta',\phi')$.
2580: In the second sum we relabel $\omega \to -\omega$ and $m \to -m$, and
2581: we use the formula (\ref{eq:Zparity1}). This gives
2582: \begin{eqnarray}
2583: \fl
2584: G_{\rm rad}(x,x') &=&
2585: \frac{1}{8 \pi i}
2586: \int_{-\infty}^\infty d\omega e^{-i \omega (t - t')}
2587: \sum_{lm} S_{\omega lm}(\theta,\phi) S_{\omega
2588: lm}(\theta',\phi')^*
2589: \frac{\omega}{|\omega|}
2590: \frac{1}{\varpi \varpi'} \nn \\
2591: \mbox{} && \times
2592: \bigg\{ \left[ u^{\rm up}_{\omega lm}(r) u^{\rm in}_{\omega lm}(r')
2593: + u^{\rm up}_{-\omega l-m}(r) u^{\rm in}_{-\omega l-m}(r') \right]
2594: \theta(r-r') \nn \\
2595: \mbox{} &&
2596: +\left[ u^{\rm in}_{\omega lm}(r) u^{\rm up}_{\omega lm}(r')
2597: + u^{\rm in}_{-\omega l-m}(r) u^{\rm up}_{-\omega l-m}(r')
2598: \right] \theta(r'-r)
2599: \bigg\}.
2600: \label{eq:Gradformula4}
2601: \end{eqnarray}
2602: Next, since our choice (\ref{eq:normalizationchoice}) of the
2603: normalization constants satisfies the conditions (\ref{eq:alphacondt}) and
2604: (\ref{eq:betacondt}), we can use the identities (\ref{eq:uinidentity})
2605: and (\ref{eq:uupidentity}).
2606: This yields
2607: \begin{eqnarray}
2608: \fl
2609: G_{\rm rad}(x,x') &=&
2610: \frac{1}{8 \pi i}
2611: \int_{-\infty}^\infty d\omega e^{-i \omega (t - t')}
2612: \sum_{lm} S_{\omega lm}(\theta,\phi) S_{\omega
2613: lm}(\theta',\phi')^*
2614: \frac{\omega}{|\omega|}
2615: \frac{1}{\varpi \varpi'} \nn \\
2616: \mbox{} && \times
2617: \bigg\{ \left[ u^{\rm up}_{\omega lm}(r) u^{\rm in}_{\omega lm}(r')
2618: + u^{\rm up}_{\omega lm}(r)^* u^{\rm in}_{\omega lm}(r')^* \right]
2619: \theta(r-r') \nn \\
2620: \mbox{} &&
2621: +\left[ u^{\rm in}_{\omega lm}(r) u^{\rm up}_{\omega lm}(r')
2622: + u^{\rm in}_{\omega lm}(r)^* u^{\rm up}_{\omega lm}(r')^*
2623: \right] \theta(r'-r)
2624: \bigg\}.
2625: \label{eq:Gradformula5}
2626: \end{eqnarray}
2627:
2628:
2629: Consider now the coefficient of $\theta(r-r')$ inside the curly
2630: brackets in Eq.\ (\ref{eq:Gradformula5}). We denote this quantity by
2631: $H(r,r')$:
2632: \be
2633: H(r,r') = u^{\rm up}(r) u^{\rm in}(r')
2634: + u^{\rm up}(r)^* u^{\rm in}(r')^* .
2635: \label{eq:Hdef}
2636: \ee
2637: Here and in the next few paragraphs we omit for simplicity the
2638: subscripts $\omega lm$. Next we use the formula
2639: (\ref{eq:uoutformula}) for the ``out'' modes in terms of the ``in''
2640: and ``up'' modes, together with our choice
2641: (\ref{eq:normalizationchoice}) of normalization constants, to obtain
2642: \be
2643: u^{\rm up}(r) = \frac{\kappa}{|\tau|^2} \left[ u^{\rm out}(r) - \sigma^*
2644: u^{\rm in}(r) \right].
2645: \label{eq:uupformula1}
2646: \ee
2647: Here
2648: \be
2649: \kappa \equiv \omega p_{m\omega} / |\omega p_{m\omega}|
2650: \ee
2651: is a variable which is $+1$ for normal modes and $-1$ for superradiant
2652: modes. Inserting this into Eq.\ (\ref{eq:Hdef}) and expanding yields
2653: \begin{eqnarray}
2654: \fl
2655: H(r,r') = \frac{\kappa}{|\tau|^2} \bigg[ u^{\rm in}(r)^* u^{\rm in}(r') +
2656: u^{\rm in}(r) u^{\rm in}(r')^*
2657: - \sigma^* u^{\rm in}(r) u^{\rm in}(r') \nonumber \\
2658: \lo - \sigma u^{\rm in}(r)^*
2659: u^{\rm in}(r')^* \bigg].
2660: \label{eq:step3}
2661: \end{eqnarray}
2662: Next, by using Eq.\ (\ref{eq:uupformula1}) to evaluate the quantity
2663: $u^{\rm up}(r)^* u^{\rm up}(r')$ we obtain the identity
2664: \begin{eqnarray}
2665: \fl
2666: |\tau|^4 u^{\rm up}(r)^* u^{\rm up}(r') - u^{\rm in}(r) u^{\rm
2667: in}(r')^* - | \sigma |^2 u^{\rm in}(r)^* u^{\rm in}(r') =
2668: - \sigma^* u^{\rm in}(r) u^{\rm in}(r') \nonumber \\
2669: \lo - \sigma u^{\rm in}(r)^* u^{\rm in}(r')^*.
2670: \end{eqnarray}
2671: We now use this identity to eliminate the last two terms inside the
2672: square brackets in Eq.\ (\ref{eq:step3}), and simplify using the
2673: identity (\ref{eq:widentity3}). This gives
2674: \be
2675: H(r,r') = u^{\rm in}(r)^* u^{\rm in}(r') + \kappa |\tau|^2 u^{\rm up}(r)^*
2676: u^{\rm up}(r').
2677: \label{eq:Hformula}
2678: \ee
2679: Now the right hand side of Eq.\ (\ref{eq:Hformula}) is explicitly
2680: invariant under the combined transformations of interchanging $r$ and
2681: $r'$ and taking the complex conjugate. However, from the definition
2682: (\ref{eq:Hdef}) of $H(r,r')$, the left hand side is invariant under
2683: complex conjugation. It follows that both sides of Eq.\
2684: (\ref{eq:Hformula}) are real and also symmetric under interchange of
2685: $r$ and $r'$:
2686: \be
2687: H(r,r') = H(r',r).
2688: \label{eq:Hsym}
2689: \ee
2690:
2691:
2692: Next, the quantity inside the curly brackets in the expression
2693: (\ref{eq:Gradformula5}) for $G_{\rm rad}$ is
2694: \be
2695: H(r,r') \theta(r-r') + H(r',r) \theta(r'-r).
2696: \ee
2697: Using the symmetry property (\ref{eq:Hsym}) together with
2698: $\theta(r-r') + \theta(r'-r) =1$, this can be written simply as
2699: $H(r,r')$. Therefore we can replace the expression in curly brackets
2700: in (\ref{eq:Gradformula5}) with the expression (\ref{eq:Hformula}) for
2701: $H(r,r')$. This gives
2702: \begin{eqnarray}
2703: \fl
2704: G_{\rm rad}(x,x') &=& \frac{1}{8 \pi i}
2705: \int_{-\infty}^\infty d\omega e^{-i \omega (t - t')}
2706: \sum_{lm} S_{\omega lm}(\theta,\phi) S_{\omega
2707: lm}(\theta',\phi')^*
2708: \frac{\omega}{|\omega|}
2709: \frac{1}{\varpi \varpi'} \nn \\
2710: \mbox{} && \times
2711: \left[
2712: u^{\rm in}_{\omega lm}(r)^* u^{\rm in}_{\omega
2713: lm}(r') +\frac{\omega p_{m\omega}}{|\omega p_{m\omega}|}
2714: |\tau_{\omega lm}|^2
2715: u^{\rm up}_{\omega lm}(r)^* u^{\rm up}_{\omega lm}(r')
2716: \right].
2717: \label{eq:Gradformula6}
2718: \end{eqnarray}
2719: Finally, rewriting this in terms of the ``down'' and ``out'' modes
2720: using the definitions (\ref{eq:uoutdef}) and (\ref{eq:udowndef}), and
2721: reinserting the appropriate factors of the normalization constants
2722: yields the formula (\ref{eq:Gradformula1}).
2723:
2724:
2725:
2726:
2727: \section{Harmonic decomposition of the scalar source}
2728: \label{sec:harmonic}
2729:
2730: \subsection{The retarded and radiative fields}
2731:
2732: Using the expression (\ref{eq:Gretformula0}) for the retarded Green's
2733: function together with the integral expression (\ref{eq:Gretdef}),
2734: we can compute the retarded field $\Phi_{\rm ret}(x)$ generated by the source
2735: ${\cal T}(x)$. For the case we are interested in, ${\cal T}(x)$ will be nonzero
2736: only in a finite range of values of $r$ of the form
2737: \be
2738: r_{\rm min} \le r \le r_{\rm max}.
2739: \ee
2740: For $r > r_{\rm max}$ only the first term in the square brackets in
2741: Eq.\ (\ref{eq:Gretformula0}) will contribute, and the function
2742: $\theta(r-r')$ will always be $1$. This gives
2743: \be
2744: \fl
2745: \Phi_{\rm ret}(x) =
2746: \frac{1}{16 \pi i} \int_{-\infty}^\infty d\omega
2747: \sum_{l=0}^\infty \sum_{m=-l}^l \frac{\omega}{|\omega|}
2748: \frac{1}{\alpha_{\omega lm} \beta_{\omega lm}} Z^{\rm out}_{\omega lm} \pi^{\rm
2749: up}_{\omega lm}(x), \ \ \ \ r \ge r_{\rm max},
2750: \ee
2751: where the coefficients $Z^{\rm out}_{\omega lm}$ are given by the inner
2752: product of the ``out'' modes with the source:
2753: \be
2754: Z^{\rm out}_{\omega lm} \equiv \left(
2755: \pi^{\rm out}_{\omega lm}, {\cal T} \right).
2756: \label{eq:Zoutdef}
2757: \ee
2758: Here following Gal'tsov \cite{Galtsov:1982} we have defined the inner
2759: product of any functions $\Phi(x)$ and
2760: $\Psi(x)$ on spacetime by
2761: \be
2762: \left( \Phi, \Psi \right) \equiv \int d^4 x \sqrt{-g(x)} \Phi(x)^*
2763: \Psi(x).
2764: \label{eq:innerproductdef}
2765: \ee
2766: Similarly for $r < r_{\rm min}$ we obtain
2767: \be
2768: \fl
2769: \Phi_{\rm ret}(x) =
2770: \frac{1}{16 \pi i} \int_{-\infty}^\infty d\omega
2771: \sum_{l=0}^\infty \sum_{m=-l}^l \frac{\omega}{|\omega|}
2772: \frac{1}{\alpha_{\omega lm} \beta_{\omega lm}} Z^{\rm down}_{\omega lm} \pi^{\rm
2773: in}_{\omega lm}(x), \ \ \ \ r \le r_{\rm min},
2774: \ee
2775: where
2776: \be
2777: Z^{\rm down}_{\omega lm} \equiv \left(
2778: \pi^{\rm down}_{\omega lm}, {\cal T} \right).
2779: \label{eq:Zdowndef}
2780: \ee
2781: Finally, the expression (\ref{eq:Gradformula0}) for the radiative
2782: Green's function together with the definitions (\ref{eq:Zoutdef}) and
2783: (\ref{eq:Zdowndef}) give the following expression for the radiative field
2784: \begin{eqnarray}
2785: \fl
2786: \Phi_{\rm rad}(x) = \frac{1}{32 \pi i} \int_{-\infty}^\infty d\omega
2787: \sum_{l=0}^\infty \sum_{m=-l}^l \frac{\omega}{|\omega|}
2788: \bigg[
2789: \frac{1}{|\alpha_{\omega lm}|^2}
2790: Z^{\rm out}_{\omega lm} \pi^{\rm out}_{\omega lm}(x)
2791: \nonumber \\
2792: \lo
2793: +\frac{\omega p_{m\omega}}{|\omega p_{m\omega}|}
2794: \frac{|\tau_{\omega lm}|^2}{|\beta_{\omega lm}|^2}
2795: Z^{\rm down}_{\omega lm} \pi^{\rm down}_{\omega lm}(x)
2796: \bigg].
2797: \label{eq:Phiradformula0}
2798: \end{eqnarray}
2799: All of these expressions depend on the amplitudes $Z^{\rm out}_{\omega
2800: lm}$ and $Z^{\rm down}_{\omega lm}$.
2801:
2802:
2803: \subsection{Harmonic decomposition of amplitudes}
2804:
2805: For the case considered here where the source ${\cal T}(x)$ is a point
2806: particle on a bound geodesic orbit, the amplitudes $Z^{\rm
2807: out}_{\omega lm}$ and $Z^{\rm down}_{\omega lm}$ can be expressed
2808: as discrete sums over delta functions \cite{Drasco:2004tv,drasconotes}:
2809: \be
2810: Z_{\omega lm}^{\rm out,down} = \sum_{k=-\infty}^\infty
2811: \sum_{n=-\infty}^\infty \, Z^{\rm out,down}_{lmkn} \, \delta(\omega - \omega_{mkn}).
2812: \label{eq:harmonicdecomposition}
2813: \ee
2814: Here
2815: \be
2816: \omega_{mkn} = m \Omega_\phi + k \Omega_\theta + n \Omega_r,
2817: \label{eq:omegamkndef}
2818: \ee
2819: \be
2820: \Omega_\phi = \frac{\Upsilon_\phi}{\Gamma},\ \ \ \
2821: \Omega_\theta = \frac{\Upsilon_\theta}{\Gamma},\ \ \ \
2822: \Omega_r = \frac{\Upsilon_r}{\Gamma},\ \ \ \
2823: \ee
2824: \be
2825: \Upsilon_\theta = \frac{2 \pi}{\Lambda_\theta},\ \ \ \
2826: \Upsilon_r = \frac{2 \pi}{\Lambda_r},\ \ \ \
2827: \label{eq:Upsilondefs}
2828: \ee
2829: and $\Gamma$ and $\Upsilon_\phi$ are defined by Eqs.\ (\ref{eq:Gammadef}) and
2830: (\ref{eq:Upsilonphidef}). The formula for the coefficients $Z^{\rm
2831: out}_{lmkn}$ and $Z^{\rm down}_{lmkn}$ is
2832: \begin{eqnarray}
2833: \fl
2834: Z^{\rm out,down}_{lmkn} = - \frac{4 \pi q}{\Gamma \Lambda_r
2835: \Lambda_\theta}
2836: e^{-i m \phi_0} e^{i \omega_{mkn} t_0}
2837: \int_0^{\Lambda_r} d\lambda_r \, \int_0^{\Lambda_\theta}
2838: d\lambda_\theta \ e^{i(k \Upsilon_\theta \lambda_\theta + n \Upsilon_r
2839: \lambda_r)} \nonumber \\
2840: \lo
2841: \times \frac{\Sigma[r(\lambda_r),\theta(\lambda_\theta)]}{\varpi[r(\lambda_r)]}
2842: \Theta_{\omega_{mkn}lm}[\theta(\lambda_\theta)]^*
2843: e^{- i m \Delta \phi_r(\lambda_r)} e^{-i m \Delta \phi_\theta(\lambda_\theta)}
2844: \nn \\ \lo \times
2845: e^{i \omega_{mkn} \Delta t_r(\lambda_r)} e^{i
2846: \omega_{mkn} \Delta t_\theta(\lambda_\theta)} u^{\rm
2847: out,down}_{\omega_{mkn}lm}[r(\lambda_r)]^*.
2848: \label{eq:Zlmknformula}
2849: \end{eqnarray}
2850: Here the functions $\Delta t_r$, $\Delta t_\theta$, $\Delta \phi_r$
2851: and $\Delta \phi_\theta$ are defined in Eqs.\
2852: (\ref{eq:deltatrdef}),
2853: (\ref{eq:deltatthetadef}),
2854: (\ref{eq:deltaphirdef}), and
2855: (\ref{eq:deltaphithetadef}). Also the function $\Theta_{\omega
2856: lm}(\theta)$ is defined by the decomposition
2857: of the spheroidal harmonic $S_{\omega lm}$:
2858: \be
2859: S_{\omega lm}(\theta,\phi) = \Theta_{\omega lm}(\theta) e^{i m \phi},
2860: \label{eq:Sdecompose}
2861: \ee
2862: cf.\ Eq.\ (\ref{eq:Somegalmdef}) above.
2863: For evaluating the double integral (\ref{eq:Zlmknformula}) it is
2864: convenient to use the identity (\ref{eq:averageidentity}) that
2865: expresses the integral in terms of the variables $\chi$ and $\psi$.
2866:
2867:
2868:
2869:
2870: In this section we review the derivation of the harmonic decomposition
2871: (\ref{eq:harmonicdecomposition}) given by Drasco and Hughes
2872: \cite{Drasco:2004tv,drasconotes}, adapted to the scalar case, and
2873: generalized from fiducial geodesics to arbitrary bound geodesics.
2874: We consider only the ``out'' amplitude $Z^{\rm
2875: out}_{\omega lm}$; the ``down'' case is exactly analogous.
2876:
2877: \subsection{Derivation}
2878:
2879:
2880: We start by inserting the definition (\ref{eq:sourcedef}) of the source
2881: ${\cal T}(x)$ into the definition of $Z_{\omega lm}^{\rm out}$ given by Eqs.\
2882: (\ref{eq:Zoutdef}) and (\ref{eq:innerproductdef}).
2883: This gives an expression consisting of an integral along the geodesic
2884: of the mode function:
2885: \be
2886: Z_{\omega lm}^{\rm out} = - q \int_{-\infty}^\infty d\tau \, \pi^{\rm
2887: out}_{\omega lm}[ t(\tau),r(\tau),\theta(\tau),\phi(\tau)]^*.
2888: \label{eq:Zlmout1}
2889: \ee
2890: Inserting the expression (\ref{eq:pioutdef}) for the mode
2891: function $\pi^{\rm out}_{\omega lm}(t,r,\theta,\phi)$ in terms of
2892: the radial mode function $u^{\rm out}_{\omega lm}(r)$, and using the
2893: decomposition (\ref{eq:Sdecompose}) of the spheroidal harmonic $S_{\omega
2894: lm}(\theta,\phi)$ now gives
2895: \be
2896: \fl
2897: Z_{\omega lm}^{\rm out} = - 2 q \int_{-\infty}^\infty d\tau \,
2898: \Theta_{\omega lm}[\theta(\tau)]^* e^{-i m \phi(\tau)} e^{i \omega
2899: t(\tau)} \frac{1}{\varpi[r(\tau)]} u_{\omega lm}^{\rm out}[r(\tau)]^*.
2900: \ee
2901: We next change the variable of integration from proper time $\tau$ to
2902: Mino time $\lambda$ using Eq.\ (\ref{eq:Minotime}), and we
2903: use the expressions (\ref{eq:tmotion00}) and (\ref{eq:phimotion00}) for
2904: the functions $t(\lambda)$ and $\phi(\lambda)$. This gives
2905: \begin{eqnarray}
2906: \fl
2907: Z_{\omega lm}^{\rm out} = - 2 q e^{-i m \phi_0} e^{ i \omega t_0}
2908: \int_{-\infty}^\infty d\lambda \,
2909: e^{i \lambda (\Gamma \omega - m \Upsilon_\phi)}
2910: \Sigma[r(\lambda),\theta(\lambda)]
2911: \Theta_{\omega lm}[\theta(\lambda)]^*
2912: \nn \\
2913: \lo
2914: \times
2915: e^{-i m \Delta \phi_r(\lambda)}
2916: e^{-i m \Delta \phi_\theta(\lambda)}
2917: e^{i \omega \Delta t_r(\lambda)}
2918: e^{i \omega \Delta t_\theta (\lambda)}
2919: \frac{1}{\varpi[r(\lambda)]} u_{\omega lm}^{\rm out}[r(\lambda)]^*.
2920: \label{eq:Zoutformula1}
2921: \end{eqnarray}
2922:
2923:
2924: We now define the function of two variables
2925: \begin{eqnarray}
2926: \fl
2927: J_{\omega lm}(\lambda_r,\lambda_\theta)
2928: &=& - 2 q e^{-i m \phi_0} e^{i \omega t_0} \
2929: \Sigma[r(\lambda_r),\theta(\lambda_\theta)]
2930: \Theta_{\omega lm}[\theta(\lambda_\theta)]^*
2931: e^{-i m \Delta \phi_r(\lambda_r)}
2932: \nn \\ \mbox{} &\times&
2933: e^{-i m \Delta \phi_\theta(\lambda_\theta)}
2934: e^{i \omega \Delta t_r(\lambda_r)}
2935: e^{i \omega \Delta t_\theta (\lambda_\theta)}
2936: \frac{1}{\varpi[r(\lambda_r)]} u_{\omega lm}^{\rm out}[r(\lambda_r)]^*.
2937: \label{eq:Jomegalmdef}
2938: \end{eqnarray}
2939: This function has two key properties. First, when evaluated at
2940: $\lambda_r = \lambda_\theta = \lambda$, it coincides with the
2941: integrand of Eq.\ (\ref{eq:Zoutformula1}), up to the first exponential factor:
2942: \be
2943: Z_{\omega lm}^{\rm out} = \int_{-\infty}^\infty d\lambda \,
2944: J_{\omega lm}(\lambda,\lambda)
2945: e^{i \lambda (\Gamma \omega - m \Upsilon_\phi)}.
2946: \label{eq:Zoutformula2}
2947: \ee
2948: Second, the function $J_{\omega lm}$ is biperiodic:
2949: \begin{eqnarray}
2950: J_{\omega lm}(\lambda_r + \Lambda_r , \lambda_\theta) &=& J_{\omega
2951: lm}(\lambda_r , \lambda_\theta), \nn \\
2952: \mbox{}
2953: J_{\omega lm}(\lambda_r , \lambda_\theta + \Lambda_\theta) &=& J_{\omega
2954: lm}(\lambda_r , \lambda_\theta).
2955: \end{eqnarray}
2956: This follows from the fact that the functions $r(\lambda)$, $\Delta
2957: t_r(\lambda)$ and $\Delta \phi_r(\lambda)$ are all periodic with
2958: period $\Lambda_r$, and the functions $\theta(\lambda)$, $\Delta
2959: t_\theta(\lambda)$ and $\Delta \phi_\theta(\lambda)$ are all periodic with
2960: period $\Lambda_\theta$, cf.\ Sec.\ \ref{sec:geosesics} above.
2961: Standard properties of biperiodic functions now imply that $J_{\omega
2962: lm}$ can be written in terms of a double Fourier series:
2963: \be
2964: J_{\omega lm}(\lambda_r,\lambda_\theta) =
2965: \sum_{k=-\infty}^\infty
2966: \sum_{n=-\infty}^\infty
2967: J_{\omega lmkn} e^{-i (k \Upsilon_\theta \lambda_\theta + n \Upsilon_r
2968: \lambda_r)},
2969: \label{eq:doubleFourierseries}
2970: \ee
2971: where $\Upsilon_\theta$ and $\Upsilon_r$ are given by Eq.\
2972: (\ref{eq:Upsilondefs}), and where the coefficients $J_{\omega lmkn}$
2973: are given by
2974: \be
2975: J_{\omega lmkn} = \frac{1}{\Lambda_r \Lambda_\theta}
2976: \int_0^{\Lambda_r} d\lambda_r
2977: \int_0^{\Lambda_\theta} d\lambda_\theta
2978: \,
2979: e^{i (k \Upsilon_\theta \lambda_\theta + n \Upsilon_r \lambda_r)}
2980: J_{\omega lm}(\lambda_r,\lambda_\theta).
2981: \label{eq:Jomegalmkndef}
2982: \ee
2983:
2984:
2985: We now insert the Fourier series (\ref{eq:doubleFourierseries}) for
2986: $J_{\omega lm}$, evaluated at $\lambda_r = \lambda_\theta = \lambda$,
2987: into the expression (\ref{eq:Zoutformula2}) for $Z^{\rm out}_{\omega lm}$.
2988: This gives
2989: \begin{eqnarray}
2990: Z^{\rm out}_{\omega lm} &=& \sum_{kn} \int_{-\infty}^\infty d\lambda
2991: e^{i \lambda ( \Gamma \omega - m \Upsilon_\phi - k \Upsilon_\theta - n
2992: \Upsilon_r) } J_{\omega lmkn} \nn \\
2993: \mbox{} &=& \sum_{kn} \frac{2 \pi}{\Gamma} \delta(\omega -
2994: \omega_{mkn}) J_{\omega lmkn},
2995: \end{eqnarray}
2996: where we have used the definition (\ref{eq:omegamkndef}) of
2997: $\omega_{mkn}$. Now comparing with Eq.\
2998: (\ref{eq:harmonicdecomposition}) we can read off that
2999: \be
3000: Z^{\rm out}_{lmkn} = \frac{2 \pi}{\Gamma} J_{\omega_{mkn} lmkn}.
3001: \label{eq:Zlmknans}
3002: \ee
3003: Combining this with the definitions (\ref{eq:Jomegalmdef}) and
3004: (\ref{eq:Jomegalmkndef}) evaluated at $\omega = \omega_{mkn}$ finally
3005: gives the formula (\ref{eq:Zlmknformula}).
3006:
3007:
3008: Note that it follows from the harmonic decomposition
3009: (\ref{eq:harmonicdecomposition}) that for geodesic sources, the
3010: continuous frequency $\omega$ and the discrete indices $l$, $m$ are
3011: replaced with the four discrete indices $k$,$n$,$l$, and $m$. In this
3012: context the operation
3013: \be
3014: \omega \to -\omega, \ \ \ \ \ m \to - m, \ \ \ \ l \to l
3015: \ee
3016: associated with the symmetries
3017: (\ref{eq:lambdaidentity}), (\ref{eq:Zparity1}), (\ref{eq:Videntity}),
3018: (\ref{eq:uinidentity}) -- (\ref{eq:tauidentity}) and
3019: (\ref{eq:uupidentity}) -- (\ref{eq:betacondt}) is replaced by the
3020: operation
3021: \be
3022: k \to -k, \ \ \ \ \ n \to -n, \ \ \ \ \ m \to -m, \ \ \ \ \ l \to l.
3023: \ee
3024:
3025:
3026: \subsection{Dependence of amplitudes on parameters of geodesic}
3027: \label{sec:dep}
3028:
3029: The amplitude $Z^{\rm out}_{lmkn}$ is a function of the parameters
3030: characterizing the geodesic, namely $E$, $L_z$, $Q$, $t_0$, $\phi_0$,
3031: $\lambda_{r0}$, and $\lambda_{\theta0}$, cf.\ Sec.\
3032: \ref{sec:geodesicparameters} above. We write this dependence as
3033: \be
3034: Z^{\rm out}_{lmkn} = Z^{\rm
3035: out}_{lmkn}(E,L_z,Q,t_0,\phi_0,\lambda_{r0},\lambda_{\theta0}).
3036: \ee
3037: We now specialize to the fiducial geodesic associated with
3038: the constants $E$, $L_z$ and $Q$, i.e. the geodesic for which
3039: $t_0=\phi_0 = \lambda_{r0} = \lambda_{\theta0} =0$.
3040: For this case we can simplify the formula
3041: (\ref{eq:Zlmknformula}) by setting $t_0$ and $\phi_0$ to zero, by
3042: replacing the motions $r(\lambda)$ and $\theta(\lambda)$ with the
3043: fiducial motions ${\hat r}(\lambda)$ and ${\hat \theta}(\lambda)$
3044: defined by Eqs.\ (\ref{eq:hatrdef}) and (\ref{eq:hatthetadef}),
3045: and by replacing the functions $\Delta t_r$, $\Delta t_\theta$,
3046: $\Delta \phi_r$ and $\Delta \phi_\theta$ with the functions ${\hat
3047: t}_r$, ${\hat t}_\theta$, ${\hat \phi}_r$, and ${\hat \phi}_\theta$.
3048: defined by Eqs.\ (\ref{eq:hattrdef}), (\ref{eq:hattthetadef}),
3049: (\ref{eq:hatphirdef}), and (\ref{eq:hatphithetadef}).
3050: This yields
3051: \begin{eqnarray}
3052: \fl
3053: Z^{\rm out}_{lmkn}(E,L_z,Q,0,0,0,0) = - \frac{4 \pi q}{\Gamma \Lambda_r
3054: \Lambda_\theta}
3055: \int_0^{\Lambda_r} d\lambda_r \, \int_0^{\Lambda_\theta}
3056: d\lambda_\theta \ e^{i(k \Upsilon_\theta \lambda_\theta + n \Upsilon_r
3057: \lambda_r)}
3058: \nn \\ \lo \times
3059: \frac{\Sigma[{\hat r}(\lambda_r),{\hat
3060: \theta}(\lambda_\theta)]}{\varpi[{\hat r}(\lambda_r)]}
3061: \Theta_{\omega_{mkn}lm}[{\hat \theta}(\lambda_\theta)]^*
3062: e^{- i m {\hat \phi}_r(\lambda_r)}
3063: \nn \\ \lo \times
3064: e^{-i m {\hat \phi}_\theta(\lambda_\theta)}
3065: e^{i \omega_{mkn} {\hat t}_r(\lambda_r)} e^{i
3066: \omega_{mkn} {\hat t}_\theta(\lambda_\theta)} u^{\rm
3067: out}_{\omega_{mkn}lm}[{\hat r}(\lambda_r)]^*.
3068: \label{eq:Zlmknformulafiducial}
3069: \end{eqnarray}
3070: For more general geodesics, the amplitude $Z_{lmkn}^{\rm out}$ depends
3071: on the parameters $t_0$,
3072: $\phi_0$, $\lambda_{r0}$ and $\lambda_{\theta0}$ only through an
3073: overall phase:
3074: \be
3075: \fl
3076: Z^{\rm out}_{lmkn}(E,L_z,Q,t_0,\phi_0,\lambda_{r0},\lambda_{\theta0})
3077: = e^{i \chi_{lmkn}(t_0,\phi_0,\lambda_{r0},\lambda_{\theta0})}
3078: Z^{\rm out}_{lmkn}(E,L_z,Q,0,0,0,0),
3079: \label{eq:changeparameter}
3080: \ee
3081: where
3082: \begin{eqnarray}
3083: \fl
3084: \chi_{lmkn}(t_0,\phi_0,\lambda_{r0},\lambda_{\theta0}) &=&
3085: k \Upsilon_\theta \lambda_{\theta0}
3086: + n \Upsilon_r \lambda_{r0}
3087: + m \left[ {\hat \phi}_r(-\lambda_{r0}) + {\hat
3088: \phi}_\theta(-\lambda_{\theta0}) - \phi_0 \right]
3089: \nn \\ \mbox{} &&
3090: -\omega_{mkn} \left[ {\hat t}_r(-\lambda_{r0}) + {\hat
3091: t}_\theta(-\lambda_{\theta0}) - t_0 \right].
3092: \label{eq:chilmnkdef}
3093: \end{eqnarray}
3094: This formula can be derived by substituting the expressions (\ref{eq:deltatrformula}),
3095: (\ref{eq:deltatthetaformula}), (\ref{eq:deltaphirformula}) and
3096: (\ref{eq:deltaphithetaformula}) for the functions $\Delta t_r$,
3097: $\Delta t_\theta$, $\Delta \phi_r$ and $\Delta \phi_\theta$ into Eq.\
3098: (\ref{eq:Zlmknformula}), making the changes of variables in the integral
3099: \be
3100: \lambda_r \to {\tilde \lambda}_r = \lambda_r - \lambda_{r0},\ \ \ \ \
3101: \lambda_\theta \to {\tilde \lambda}_\theta = \lambda_\theta -
3102: \lambda_{\theta0},
3103: \ee
3104: and comparing with Eq.\ (\ref{eq:Zlmknformulafiducial}). Finally we
3105: note that the phase (\ref{eq:chilmnkdef}) and amplitude
3106: (\ref{eq:changeparameter}) are invariant under the transformations
3107: (\ref{eq:t1}), (\ref{eq:t2}), (\ref{eq:t3}), and (\ref{eq:t4}) that
3108: correspond to the re-parameterization $\lambda \to \lambda + \Delta
3109: \lambda$. This invariance serves as a consistency check of the
3110: formulae, since we expect the invariance on physical grounds.
3111:
3112:
3113:
3114:
3115: \section{Expressions for the time derivatives of energy, angular
3116: momentum, and rest mass}
3117: \label{sec:Edot}
3118:
3119: \subsection{Time averages}
3120:
3121: Let ${\cal E}$ be one of the three conserved quantities of geodesic
3122: motion, $E$, $L_z$ or $Q$. For the purpose of evolving the orbit we
3123: would like to compute the quantity
3124: \be
3125: \left< \frac{d{\cal E} }{d t} \right>_t,
3126: \label{eq:desired}
3127: \ee
3128: that is, the average with respect to the Boyer-Lindquist time
3129: coordinate $t$ of the derivative of ${\cal E}$ with respect to $t$.
3130: However, the quantity that is most naturally computed is the
3131: derivative with respect to proper time $\tau$, and the type of average
3132: that is most easily computed is the average with respect to Mino time
3133: $\lambda$. In this section we therefore rewrite the quantity
3134: (\ref{eq:desired}) in terms of a Mino-time average of $d{\cal
3135: E}/d\tau$.
3136:
3137: In the adiabatic limit, we can choose a time interval $\Delta t$ which
3138: is long compared to the orbital timescales but short compared to the
3139: radiation reaction time\footnote{A natural choice for $\Delta t$ is the geometric
3140: mean of the orbital time and the radiation reaction time; this is
3141: the time it takes for the phase difference between the geodesic orbit
3142: and the true orbit to become of order unity.}.
3143: Then, to a good approximation we have
3144: \be
3145: \left< \frac{d {\cal E}} {d t} \right>_t = \frac{\Delta {\cal E}}{\Delta t},
3146: \ee
3147: where $\Delta {\cal E}$ is the change in ${\cal E}$ over this
3148: interval. Now let $\Delta \lambda$ be the change in Mino time over
3149: the interval. From Eq.\ (\ref{eq:tmotion1}) we have
3150: \be
3151: \Delta t = \Gamma \Delta \lambda + {\rm oscillatory~terms}.
3152: \ee
3153: Now the oscillatory terms will be bounded as $\Delta t$ is taken
3154: larger and larger, and therefore in the adiabatic limit they will
3155: give a negligible fractional correction to $\Delta t$. Hence we get
3156: \begin{eqnarray}
3157: \left< \frac{d {\cal E}} {d t} \right>_t &=& \frac{1}{\Gamma}
3158: \frac{\Delta {\cal E}}{\Delta \lambda} \nn \\
3159: \mbox{} &=&
3160: \frac{1}{\Gamma} \left< \frac{d {\cal E}} {d \lambda} \right>_\lambda,
3161: \label{eq:dcalEdt00}
3162: \end{eqnarray}
3163: where the $\lambda$ subscript on the angular brackets means an average
3164: with respect to $\lambda$.
3165: Note that using the definition (\ref{eq:Gammadef}) of $\Gamma$ we
3166: can rewrite this formula as
3167: \begin{equation}
3168: \left< \frac{d {\cal E}} {d t} \right>_t =
3169: \frac{ \left< d {\cal E} / d \lambda \right>_\lambda }
3170: {\left< d t / d \lambda \right>_\lambda }.
3171: \end{equation}
3172: Finally we can use Eq.\ (\ref{eq:Minotime}) to rewrite the Mino-time
3173: derivative in Eq.\ (\ref{eq:dcalEdt00}) in
3174: terms of a proper time derivative. This gives the final
3175: formula which we will use:
3176: \begin{eqnarray}
3177: \left< \frac{d {\cal E}} {d t} \right>_t &=&
3178: %\frac{1}{\Gamma}
3179: %\frac{\Delta {\cal E}}{\Delta \lambda} \nn \\
3180: %\mbox{} &=&
3181: \frac{1}{\Gamma} \left< \Sigma \, \frac{d {\cal E}} {d \tau} \right>_\lambda.
3182: \label{eq:used}
3183: \end{eqnarray}
3184:
3185:
3186:
3187: \subsection{Formulae for time derivatives of energy and angular momentum}
3188:
3189: We now specialize to the cases of energy and angular momentum, ${\cal
3190: E} = E$ or ${\cal E} = L_z$. For these cases the result for the
3191: time-averaged time derivative is
3192: \begin{eqnarray}
3193: \fl
3194: \left< \frac{d {\cal E}}{dt} \right>_t = - \frac{1}{64 \pi^2 \mu}
3195: \sum_\Lambda
3196: \left\{ \begin{array}{l} \omega_{mkn} \\
3197: m \\
3198: \end{array} \right\}
3199: \frac{\omega_{mkn}}{|\omega_{mkn}|}
3200: \left[
3201: \frac{1}{|\alpha_\Lambda|^2}
3202: |Z^{\rm out}_\Lambda |^2
3203: +\frac{\omega_{mkn} p_{mkn}}{|\omega_{mkn} p_{mkn}|}
3204: \frac{|\tau_\Lambda|^2}{|\beta_\Lambda|^2}
3205: |Z^{\rm down}_\Lambda |^2 \right]. \nn \\
3206: \label{eq:dcalEdt2}
3207: \end{eqnarray}
3208: Here
3209: \be
3210: p_{mkn} \equiv p_{m\omega_{mkn}} = \omega_{mkn} - m \omega_+,
3211: \ee
3212: cf.\ Eq.\ (\ref{eq:kdef}) above.
3213: Also the symbol $\Lambda$ is a shorthand for the set of indices
3214: $lmkn$; we define
3215: \be
3216: \alpha_\Lambda = \alpha_{lmkn} \equiv \alpha_{\omega_{mkn}lm},
3217: \label{eq:alphaLambdadef}
3218: \ee
3219: and similarly for $\beta_\Lambda$ and $\tau_\Lambda$. The sum over
3220: $\Lambda$ is defined as
3221: \be
3222: \sum_\Lambda = \sum_{k=-\infty}^\infty \sum_{n=-\infty}^\infty
3223: \sum_{l=0}^\infty \sum_{m=-l}^l.
3224: \ee
3225: The expression in curly brackets in Eq.\ (\ref{eq:dcalEdt2})
3226: means either a factor of $\omega_{mkn}$ (for energy) or a factor of
3227: $m$ (for angular momentum).
3228: The first term in the square brackets corresponds to the flux through
3229: future null infinity, and the second term corresponds the flux through
3230: the horizon. Note that for energy the future null infinity term is always
3231: negative; this makes sense since the expression (\ref{eq:dcalEdt2})
3232: is the rate of change of orbital energy.
3233:
3234:
3235: The flux expression (\ref{eq:dcalEdt2})
3236: has been derived from the radiation
3237: reaction force. As shown by Gal'tsov, the same result
3238: is obtained by computing the flux directly using the stress-energy
3239: tensor of the scalar field \cite{Galtsov:1982}.
3240: %[A special case of this
3241: %result has been independently derived in Ref.\ \cite{Gralla:2005et}]
3242:
3243:
3244: Note that the final answer (\ref{eq:dcalEdt2}) is independent of
3245: our choices for the normalization constants $\alpha_\Lambda$ and
3246: $\beta_\Lambda$, since the amplitudes $Z^{\rm out}_\Lambda$ and
3247: $Z^{\rm down}_\Lambda$ contain factors of $\alpha_\Lambda$ and
3248: $\beta_\Lambda$ that compensate for the factors that appear explicitly
3249: in Eq.\ (\ref{eq:dcalEdt2}).
3250:
3251:
3252: \subsection{Derivation}
3253: \label{sec:derivation1}
3254:
3255: For energy and angular momentum, we can write the conserved quantity as
3256: the inner product of a Killing vector $\xi^\alpha$ with the 4-velocity
3257: of the particle:
3258: \be
3259: {\cal E} = \xi_\alpha u^\alpha.
3260: \ee
3261: Here ${\vec \xi} = - \partial/\partial t$ for ${\cal E} = E$, and
3262: ${\vec \xi} = \partial /\partial \phi$ for ${\cal E} = L_z$.
3263: Taking a time derivative gives
3264: \be
3265: \frac{d {\cal E}}{d \tau} = u^\beta \nabla_\beta ( \xi_\alpha
3266: u^\alpha) = (\nabla_\beta \xi_\alpha) u^\beta u^\alpha + \xi_\alpha
3267: (u^\beta \nabla_\beta u^\alpha) = \xi_\alpha a^\alpha,
3268: \ee
3269: where $a^\alpha$ is the 4-acceleration. Here we have used the fact
3270: that ${\vec \xi}$ is a Killing vector so $\nabla_{(\alpha}
3271: \xi_{\beta)} =0$.
3272:
3273: Next, we use the formula (\ref{eq:selfacc0}) for the
3274: self-acceleration. This gives
3275: \begin{eqnarray}
3276: \frac{d {\cal E}}{d \tau} &=& \frac{q}{\mu} \left[ \xi^\alpha
3277: \nabla_\alpha \Phi_{\rm rad} + (\xi_\alpha u^\alpha) u^\beta
3278: \nabla_\beta \Phi_{\rm rad} \right] \nn \\
3279: \mbox{} &=& \frac{q}{\mu} \xi^\alpha
3280: \nabla_\alpha \Phi_{\rm rad} +\frac{q}{\mu} {\cal E} \frac{d \Phi_{\rm
3281: rad}}{d \tau}.
3282: \label{eq:dcalEdtau0}
3283: \end{eqnarray}
3284: Now the second term here is a total time derivative to leading order
3285: in $q$, since ${\cal E}$ is conserved to zeroth order in $q$. Hence
3286: the change in ${\cal E}$ over an interval from $\tau_1$ to $\tau_2$
3287: associated with the second term will be
3288: \be
3289: \frac{q}{\mu} {\cal E} \left[ \Phi_{\rm rad}(\tau_2) - \Phi_{\rm
3290: rad}(\tau_1) \right].
3291: \ee
3292: This is a term which will oscillate but will not grow secularly with
3293: time (since $\Phi_{\rm rad}$ does not grow secularly with time).
3294: Hence in the adiabatic limit this term will be smaller
3295: than the change due to the first term in Eq.\
3296: (\ref{eq:dcalEdtau0}) by the ratio of the orbital timescale to the
3297: inspiral timescale, which is negligible in the approximation we are using.
3298: Dropping the second term in Eq.\
3299: (\ref{eq:dcalEdtau0}) and substituting
3300: into Eq.\ (\ref{eq:used}) we get
3301: \be
3302: \left< \frac{d {\cal E}} {d t} \right>_t = \frac{q}{\mu \Gamma}
3303: \, \left< \,\Sigma \, \xi^\alpha \nabla_\alpha \Phi_{\rm rad} \right>_\lambda.
3304: \label{eq:dcalEdtau1}
3305: \ee
3306:
3307:
3308: Next, we can obtain an explicit expression for the radiative field
3309: $\Phi_{\rm rad}$ by substituting the harmonic decomposition
3310: (\ref{eq:harmonicdecomposition}) into Eq.\ (\ref{eq:Phiradformula0}).
3311: This gives
3312: \begin{eqnarray}
3313: \fl
3314: \Phi_{\rm rad}(x) = \frac{1}{32 \pi i} \sum_{k = -\infty}^\infty
3315: \sum_{n=-\infty}^\infty \sum_{l=0}^\infty \sum_{m=-l}^l
3316: \frac{\omega_{mkn}}{|\omega_{mkn}|}
3317: \bigg[
3318: \frac{1}{|\alpha_\Lambda|^2}
3319: Z^{\rm out}_\Lambda \pi^{\rm out}_\Lambda(x)
3320: \nonumber \\
3321: \lo
3322: +\frac{\omega_{mkn} p_{mkn}}{|\omega_{mkn} p_{mkn}|}
3323: \frac{|\tau_\Lambda|^2}{|\beta_\Lambda|^2}
3324: Z^{\rm down}_\Lambda \pi^{\rm down}_\Lambda(x)
3325: \bigg].
3326: \label{eq:Phiradformula0a}
3327: \end{eqnarray}
3328: Here as above $\Lambda$ denotes the set of indices $lmkn$, and we have
3329: defined
3330: \be
3331: \pi^{\rm out,down}_{lmkn} = \pi^{\rm out,down}_{\omega_{mkn}lm}.
3332: \ee
3333: We now substitute this expression into Eq.\
3334: (\ref{eq:dcalEdtau1}), and use the fact that the operator $\xi^\alpha
3335: \nabla_\alpha$ gives a factor of $i \omega$ (for the energy) or $i m$
3336: (for the angular momentum). We get
3337: \begin{eqnarray}
3338: \fl
3339: \left< \frac{d {\cal E}}{dt} \right>_t = \frac{q}{32 \pi \mu \Gamma}
3340: \sum_\Lambda
3341: \left\{ \begin{array}{l} \omega_{mkn} \\
3342: m \\
3343: \end{array} \right\}
3344: \frac{\omega_{mkn}}{|\omega_{mkn}|}
3345: \bigg[
3346: \frac{1}{|\alpha_\Lambda|^2}
3347: Z^{\rm out}_\Lambda \left< \Sigma \, \pi^{\rm out}_\Lambda \right>_\lambda
3348: \nonumber \\
3349: \lo
3350: +\frac{\omega_{mkn} p_{mkn}}{|\omega_{mkn} p_{mkn}|}
3351: \frac{|\tau_\Lambda|^2}{|\beta_\Lambda|^2}
3352: Z^{\rm down}_\Lambda \left< \Sigma \, \pi^{\rm down}_\Lambda \right>_\lambda
3353: \bigg],
3354: \label{eq:dcalEdt1}
3355: \end{eqnarray}
3356: where the expression in curly brackets means either a factor of
3357: $\omega_{mkn}$ or a factor of $m$.
3358:
3359:
3360: We now discuss the evaluation of the quantity
3361: \be
3362: \left< \Sigma \, \pi_{\omega_{mkn}lm}^{\rm out} \right>_\lambda
3363: \ee
3364: which appears in Eq.\ (\ref{eq:dcalEdt1}). This is the time average
3365: of the ``out'' mode function evaluated on the geodesic, and can be easily
3366: evaluated using the harmonic decomposition of Sec.\
3367: \ref{sec:harmonic}. By comparing the integrands of Eqs.\
3368: (\ref{eq:Zlmout1}) and (\ref{eq:Zoutformula2}) we find that
3369: \be
3370: \fl
3371: \Sigma[z^\alpha(\lambda)] \pi_{\omega_{mkn}lm}^{\rm
3372: out}[z^\alpha(\lambda)] = - \frac{1}{q}
3373: J_{\omega_{mkn}lm}(\lambda,\lambda)^* e^{-i
3374: \lambda (\Gamma \omega_{mkn} - \Upsilon_\phi m)},
3375: \ee
3376: where $z^\alpha(\lambda) = [ t(\lambda), r(\lambda),
3377: \theta(\lambda),\phi(\lambda)]$ is the geodesic.
3378: Now using the Fourier series expansion (\ref{eq:doubleFourierseries})
3379: of $J_{\omega lm}$ gives
3380: \be
3381: \fl
3382: \Sigma[z^\alpha(\lambda)] \pi_{\omega_{mkn}lm}^{\rm
3383: out}[z^\alpha(\lambda)] = - \frac{1}{q} \sum_{k',n'}
3384: J_{\omega_{mkn}lmk'n'}^*e^{-i
3385: \lambda \Gamma (\omega_{mkn} - \omega_{mk'n'})}.
3386: \ee
3387: From this expression it is clear that averaging over $\lambda$ kills
3388: all the terms in the sum except $k'=k$, $n'=n$, and we obtain
3389: \be
3390: \fl
3391: \left< \Sigma[z^\alpha(\lambda)] \pi_{\omega_{mkn}lm}^{\rm
3392: out}[z^\alpha(\lambda)] \right>_\lambda = - \frac{1}{q}
3393: J_{\omega_{mkn}lmkn}^* = - \frac{\Gamma}{2 \pi q} (Z^{\rm out}_{lmkn})^*.
3394: \label{eq:Zoutuseful}
3395: \ee
3396: For the last equation we have used Eq.\ (\ref{eq:Zlmknans}).
3397: Finally substituting this result, together with a similar equation for
3398: the ``down'' modes, in to Eq.\ (\ref{eq:dcalEdt1}) we obtain the final
3399: result (\ref{eq:dcalEdt2})
3400:
3401:
3402: \subsection{Time derivative of renormalized rest mass}
3403:
3404: Equation (\ref{eq:dmudt2}) for the time derivative of the renormalized rest
3405: mass $\mu(\tau)$ can be immediately integrated with respect to $\tau$ between two
3406: times $\tau_1$ and $\tau_2$. This yields
3407: \be
3408: \mu(\tau_2) - \mu(\tau_1) = q \Phi_{\rm rad}[z(\tau_2)] - q \Phi_{\rm
3409: rad}[z(\tau_1)],
3410: \ee
3411: where $x^\alpha = z^\alpha(\tau)$ is the geodesic.
3412: Now since the zeroth-order orbit is bound, the right hand side of this
3413: equation does not contain any secularly growing terms. Instead it
3414: consists only of oscillatory terms. Therefore, over long timescales
3415: and in the adiabatic limit (the regime in which we are working in this
3416: paper), the rest mass is conserved:
3417: \be
3418: \frac{d\mu}{d\tau} =0.
3419: \ee
3420:
3421:
3422:
3423: \section{Expression for the time derivative of the Carter constant}
3424: \label{sec:Kdot}
3425:
3426: \subsection{Formula for time derivative}
3427:
3428:
3429: We now turn to the corresponding computation for the Carter constant
3430: $K$ defined by Eq.\ (\ref{eq:Kdef}). The result is
3431: \begin{eqnarray}
3432: \fl
3433: \left< \frac{dK}{dt} \right>_t = - \frac{1}{32 \pi^2 \mu} \sum_\Lambda
3434: \frac{\omega_{mkn}}{|\omega_{mkn}|}
3435: \left[
3436: \frac{1}{|\alpha_\Lambda|^2}
3437: ( {\tilde Z}^{\rm out}_\Lambda)^*
3438: Z^{\rm out}_\Lambda
3439: +\frac{\omega_{mkn} p_{mkn}}{|\omega_{mkn} p_{mkn}|}
3440: \frac{|\tau_\Lambda|^2}{|\beta_\Lambda|^2}
3441: ({\tilde Z}^{\rm down}_\Lambda)^* Z^{\rm down}_\Lambda
3442: \right].
3443: \nonumber \\
3444: \label{eq:dKdt}
3445: \end{eqnarray}
3446: This expression has the same structure as the expression
3447: (\ref{eq:dcalEdt2}) for
3448: the time derivatives of the energy and angular momentum, except that
3449: the squared amplitudes $|Z^{\rm out}_\Lambda|^2$ and $|Z^{\rm
3450: down}_\Lambda|^2$ have been replaced with products of the amplitudes
3451: $Z^{\rm out}_\Lambda$ and $Z^{\rm down}_\Lambda$ with two new
3452: amplitudes
3453: ${\tilde Z}^{\rm out}_\Lambda$ and ${\tilde Z}^{\rm down}_\Lambda$.
3454: These new amplitudes are defined by the equation
3455: \begin{eqnarray}
3456: \fl
3457: {\tilde Z}^{\rm out,down}_{lmkn} = \frac{4 \pi q}{\Gamma \Lambda_r
3458: \Lambda_\theta}
3459: e^{-i m \phi_0} e^{i \omega_{mkn} t_0}
3460: \int_0^{\Lambda_r} d\lambda_r \, \int_0^{\Lambda_\theta}
3461: d\lambda_\theta \ e^{i(k \Upsilon_\theta \lambda_\theta + n \Upsilon_r
3462: \lambda_r)}
3463: \Theta_{\omega_{mkn}lm}[\theta(\lambda_\theta)]^*
3464: \nn \\ \mbox{} \times
3465: e^{- i m \Delta \phi_r(\lambda_r)} e^{-i m \Delta \phi_\theta(\lambda_\theta)}
3466: e^{i \omega_{mkn} \Delta t_r(\lambda_r)} e^{i
3467: \omega_{mkn} \Delta t_\theta(\lambda_\theta)}
3468: \nn \\ \mbox{}
3469: \times \left\{ G_{mkn}[\lambda_r,\lambda_\theta]
3470: R^{\rm out,down}_{\omega_{mkn}lm}[r(\lambda_r)]^*
3471: + G[\lambda_r,\lambda_\theta]
3472: \frac{d R^{\rm out,down}_{\omega_{mkn}lm}}{dr}[r(\lambda_r)]^*
3473: \right\},
3474: \nonumber \\
3475: \label{eq:tildeZlmknformula}
3476: \end{eqnarray}
3477: where
3478: \be
3479: R^{\rm out,down}_{\omega_{mkn}lm}(r) = \frac{u^{\rm
3480: out,down}_{\omega_{mkn}lm}(r)}{\varpi(r)},
3481: \ee
3482: \be
3483: G_{mkn}(\lambda_r,\lambda_\theta) \equiv -
3484: \frac{\Sigma}{\Delta} (\varpi^2 E - a L_z)
3485: (\varpi^2 \omega_{mkn}- a m) +
3486: %\omega_{mkn} r^2 T,
3487: 2 i r u_r \Delta,
3488: \label{eq:Gmnkdef}
3489: \ee
3490: \be
3491: G(\lambda_r,\lambda_\theta) = - i \Delta \Sigma u_r.
3492: \label{eq:Gdef1}
3493: \ee
3494: On the right hand sides of Eqs.\ (\ref{eq:Gmnkdef}) and
3495: (\ref{eq:Gdef1}), it is understood that $\Sigma(r,\theta)$,
3496: $\Delta(r)$ and $\varpi(r)$ are evaluated at $r = r(\lambda_r)$ and
3497: $\theta = \theta(\lambda_\theta)$. Also it is understood that
3498: $u_r$ is given as a function of $r$ by Eq.\
3499: (\ref{eq:urformula}), and hence as a function of $\lambda_r$ by using $r
3500: = r(\lambda_r)$ and by resolving the sign ambiguity in
3501: Eq.\ (\ref{eq:urformula}) using Eq.\ (\ref{eq:psidef}).
3502:
3503:
3504: Several features of the formulae (\ref{eq:dKdt}) and
3505: (\ref{eq:tildeZlmknformula}) are worth noting:
3506:
3507:
3508:
3509: \begin{itemize}
3510:
3511: \item The formula (\ref{eq:tildeZlmknformula}) for the new amplitudes
3512: ${\tilde Z}^{\rm out}_{lmkn}$ and ${\tilde Z}^{\rm down}_{lmkn}$.
3513: has a very similar structure to the formula (\ref{eq:Zlmknformula})
3514: defining the amplitudes $Z^{\rm out}_{lmkn}$ and $Z^{\rm
3515: down}_{lmkn}$. In particular, the derivation of the formula
3516: (\ref{eq:changeparameter}) for the dependence of the amplitudes on
3517: the parameters $t_0$, $\phi_0$, $\lambda_{r0}$ and $\lambda_{\theta0}$
3518: of the geodesic (via an overall phase) carries through as before.
3519: This means that the time derivative of the Carter constant is
3520: independent of these parameters, as expected, since the phase from the untilded
3521: amplitudes cancels the phase from the tilded amplitudes in Eq.\
3522: (\ref{eq:dKdt}).
3523:
3524: \item Although we have no general definition for ``flux of Carter'',
3525: presumably the first term in Eq.\ (\ref{eq:dKdt}) corresponds to
3526: something like the flux of Carter to future null infinity, and the
3527: second term to the flux of Carter down the black hole horizon.
3528:
3529: \item As was the case for the original amplitudes, to
3530: evaluate the double integral (\ref{eq:tildeZlmknformula}) it is
3531: convenient to use the identity (\ref{eq:averageidentity}) to
3532: express the integral in terms of the variables $\chi$ and $\psi$
3533: instead of $\lambda_\theta$ and $\lambda_r$.
3534:
3535: \end{itemize}
3536:
3537:
3538:
3539: \subsection{Derivation}
3540:
3541:
3542: Taking a time derivative of the expression (\ref{eq:Kdef}) for $K$ gives
3543: \begin{eqnarray}
3544: \frac{d K}{d \tau} &=& u^\gamma \nabla_\gamma (K_{\alpha\beta} u^\alpha u^\beta) =
3545: \nabla_{(\gamma} K_{\alpha\beta)} u^\gamma u^\alpha u^\beta + 2 K_{\alpha\beta} u^\alpha u^\gamma \nabla_\gamma u^\beta \nn \\
3546: \mbox{} &=& 2 K_{\alpha\beta} u^\alpha a^\beta.
3547: \end{eqnarray}
3548: Here we have used the Killing tensor equation $\nabla_{(\gamma} K_{\alpha\beta)} =0$.
3549: Now substituting the expression (\ref{eq:selfacc0}) for the
3550: self-acceleration gives
3551: \begin{eqnarray}
3552: \frac{d K}{d \tau} &=& \frac{2 q}{\mu} \left[ K_{\alpha\beta} u^\alpha
3553: u^\beta u^\gamma \nabla_\gamma \Phi_{\rm rad} + K_{\alpha\beta}
3554: u^\alpha \nabla^\beta \Phi_{\rm rad} \right]
3555: \nn \\
3556: \mbox{} &=& \frac{2 q}{\mu} K \frac{d \Phi_{\rm rad}}{d \tau}
3557: +\frac{2 q}{\mu} K^{\alpha\beta} u_{\alpha} \nabla_\beta \Phi_{\rm rad}.
3558: \label{eq:dKdtau0}
3559: \end{eqnarray}
3560: Now the first term here is a total time derivative to leading order
3561: in $q$, since $K$ is conserved to zeroth order in $q$. Hence
3562: this term can be neglected for the reason explained in Sec.\
3563: \ref{sec:derivation1}. Dropping the first term and substituting
3564: into Eq.\ (\ref{eq:used}) we get
3565: \be
3566: \left< \frac{d K} {d t} \right>_t = \frac{2 q}{\mu \Gamma}
3567: \, \left< \,\Sigma \, K^{\alpha\beta} u_\alpha \nabla_\beta \Phi_{\rm rad} \right>_\lambda.
3568: \label{eq:dKdtau1}
3569: \ee
3570:
3571:
3572: Next, we use the explicit expression (\ref{eq:Phiradformula0a})
3573: for the radiation field $\Phi_{\rm rad}$.
3574: This gives
3575: \begin{eqnarray}
3576: \fl
3577: \left< \frac{d K}{dt} \right>_t = \frac{q}{16 \pi i \mu \Gamma}
3578: \sum_\Lambda
3579: \frac{\omega_{mkn}}{|\omega_{mkn}|}
3580: \bigg[
3581: \frac{1}{|\alpha_\Lambda|^2}
3582: Z^{\rm out}_\Lambda \left< \Sigma K^{\alpha\beta} u_\alpha
3583: \nabla_\beta \, \pi^{\rm out}_\Lambda \right>_\lambda
3584: \nonumber \\
3585: \lo
3586: +\frac{\omega_{mkn} p_{mkn}}{|\omega_{mkn} p_{mkn}|}
3587: \frac{|\tau_\Lambda|^2}{|\beta_\Lambda|^2}
3588: Z^{\rm down}_\Lambda \left< \Sigma K^{\alpha\beta} u_\alpha
3589: \nabla_\beta \, \pi^{\rm down}_\Lambda \right>_\lambda
3590: \bigg].
3591: \label{eq:dKdt1}
3592: \end{eqnarray}
3593: We now define the amplitudes
3594: \be
3595: {\tilde Z}^{\rm out}_{lmkn} = \frac{2 \pi q}{\Gamma i} \left< \Sigma
3596: K^{\alpha\beta} u_\alpha \nabla_\beta ( \pi^{\rm out}_{\omega_{mkn}lm}
3597: )^* \right>_\lambda
3598: \label{eq:tildeZoutdef}
3599: \ee
3600: and
3601: \be
3602: {\tilde Z}^{\rm down}_{lmkn} = \frac{2 \pi q}{\Gamma i} \left< \Sigma
3603: K^{\alpha\beta} u_\alpha \nabla_\beta ( \pi^{\rm down}_{\omega_{mkn}lm}
3604: )^* \right>_\lambda.
3605: \label{eq:tildeZdowndef}
3606: \ee
3607: Substituting these definitions into Eq.\ (\ref{eq:dKdt1}) gives the
3608: formula (\ref{eq:dKdt}). Therefore it remains only to derive the
3609: formula (\ref{eq:tildeZlmknformula}) for the amplitudes ${\tilde
3610: Z}_\Lambda^{\rm out}$ and ${\tilde Z}_\Lambda^{\rm down}$.
3611: We will derive the formula for
3612: ${\tilde Z}_\Lambda^{\rm out}$; the derivation of the formula for
3613: ${\tilde Z}_\Lambda^{\rm down}$ is similar.
3614:
3615:
3616: We start by simplifying the differential operator $K^{\alpha\beta} u_\alpha
3617: \nabla_\beta$ which appears in the definitions (\ref{eq:tildeZoutdef})
3618: and (\ref{eq:tildeZdowndef}). Using the expression
3619: (\ref{eq:Kalphabeta}) for the Killing tensor we can write this as
3620: \be
3621: K^{\alpha\beta} u_\alpha \nabla_\beta
3622: = \Sigma (l^\alpha u_\alpha) n^\beta \nabla_\beta
3623: + \Sigma (n^\alpha u_\alpha) l^\beta \nabla_\beta + r^2
3624: \frac{d}{d\tau}.
3625: \ee
3626: Using the definitions
3627: (\ref{eq:vecldef}) and (\ref{eq:vecndef}) of ${\vec l}$ and ${\vec
3628: n}$, the definitions (\ref{eq:Edef}) and (\ref{eq:Lzdef}) of $E$ and
3629: $L_z$, and the fact that $\partial_t$ and $\partial_\phi$ reduce to $i
3630: \omega_{mkn}$ and $-i m$ when acting on $(\pi^{\rm
3631: out}_{\omega_{mkn}lm})^*$ gives
3632: \begin{eqnarray}
3633: \fl
3634: K^{\alpha\beta} u_\alpha \nabla_\beta
3635: &=& r^2 \frac{d}{d\tau} - \frac{i}{2 \Delta}( -\varpi^2 E + a L_z +
3636: u_r \Delta )
3637: (-\varpi^2 \omega_{mkn} + a m - i \Delta \partial_r)
3638: \nn \\ \mbox{} &&
3639: - \frac{i}{2 \Delta}( -\varpi^2 E + a L_z - u_r \Delta )
3640: (-\varpi^2 \omega_{mkn} + a m + i \Delta \partial_r) \\
3641: \mbox{} &=& r^2 \frac{d}{d\tau} - \frac{i}{\Delta}( -\varpi^2 E + a
3642: L_z) (-\varpi^2 \omega_{mkn} + a m) - \Delta u_r \partial_r.
3643: \label{eq:K3}
3644: \end{eqnarray}
3645:
3646: Consider now the contribution of the first term in Eq.\
3647: (\ref{eq:K3}) to the expression (\ref{eq:tildeZoutdef})
3648: for ${\tilde Z}^{\rm out}_{lmkn}$. Using the relation (\ref{eq:Minotime})
3649: between $d\tau$ and $d\lambda$ we can write this as
3650: \begin{equation}
3651: \frac{2 \pi q}{\Gamma i} \left< \Sigma r^2 \frac{d}{d\tau}
3652: ( \pi^{\rm out}_{\omega_{mkn}lm}
3653: )^* \right>_\lambda =
3654: \frac{2 \pi q}{\Gamma i} \left< r^2 \frac{d}{d\lambda}
3655: ( \pi^{\rm out}_{\omega_{mkn}lm}
3656: )^* \right>_\lambda.
3657: \end{equation}
3658: We can integrate this by parts with respect to $\lambda$; the boundary
3659: term which is generated can be neglected for the reason explained in
3660: Sec.\ \ref{sec:derivation1}. This gives
3661: \begin{equation}
3662: - \frac{2 \pi q}{\Gamma i} \left< \frac{d r^2}{d\lambda}
3663: ( \pi^{\rm out}_{\omega_{mkn}lm}
3664: )^* \right>_\lambda =
3665: \frac{2 \pi q}{\Gamma} \left< 2 i r u_r \Delta
3666: ( \pi^{\rm out}_{\omega_{mkn}lm}
3667: )^* \right>_\lambda,
3668: \end{equation}
3669: where we have used $dr/d\lambda = \Sigma u^r = \Delta u_r$ from Eqs.\
3670: (\ref{eq:Minotime}) and (\ref{eq:kerrmetric}).
3671: Combining this with the result obtained from substituting the
3672: second and third terms in Eq.\ (\ref{eq:K3}) into
3673: Eq. (\ref{eq:tildeZoutdef}) gives
3674: \be
3675: \fl
3676: {\tilde Z}^{\rm out}_{lmkn} = \frac{2 \pi q}{\Gamma } \left<
3677: G_{mkn}(\lambda,\lambda) ( \pi^{\rm out}_{\omega_{mkn}lm} )^*
3678: + G(\lambda,\lambda) \partial_r ( \pi^{\rm out}_{\omega_{mkn}lm} )^*
3679: \right>_\lambda,
3680: \label{eq:tildeZoutformula1}
3681: \ee
3682: where the functions $G_{mkn}(\lambda_r,\lambda_\theta)$ and
3683: $G(\lambda_r,\lambda_\theta)$ are
3684: [cf.\ Eqs.\ (\ref{eq:Gmnkdef}) and (\ref{eq:Gdef1}) above]
3685: \be
3686: G_{mkn}(\lambda_r,\lambda_\theta) \equiv - \frac{\Sigma}{\Delta}
3687: (\varpi^2 E -a L_z)
3688: (\varpi^2 \omega_{mkn} - a m) + 2 i r u_r \Delta,
3689: \label{eq:Gmnkdef2}
3690: \ee
3691: \be
3692: G(\lambda_r,\lambda_\theta) = - i \Delta \Sigma u_r.
3693: \label{eq:Gdef2}
3694: \ee
3695: As explained above, it is understood that the functions $\Sigma(r,\theta)$,
3696: $\Delta(r)$ and $\varpi(r)$ on the right hand sides of Eqs.\
3697: (\ref{eq:Gmnkdef2}) and (\ref{eq:Gdef2})
3698: are evaluated at $r = r(\lambda_r)$ and
3699: $\theta = \theta(\lambda_\theta)$. Also it is understood that
3700: $u_r$ is given as a function of $r$ by Eq.\
3701: (\ref{eq:urformula}), and hence as a function of $\lambda_r$ by using $r
3702: = r(\lambda_r)$ and by resolving the sign ambiguity in
3703: Eq.\ (\ref{eq:urformula}) using Eq.\ (\ref{eq:psidef}).
3704:
3705:
3706: The average over $\lambda$ in Eq.\ (\ref{eq:tildeZoutformula1}) can be
3707: evaluated using the same techniques as in Secs.\ \ref{sec:harmonic}
3708: and \ref{sec:Edot} above. Using the definition (\ref{eq:pioutdef}) of the
3709: ``out'' mode function and the definition (\ref{eq:omegamkndef}) of
3710: $\omega_{mkn}$,
3711: the quantity inside the angular brackets in
3712: Eq.\ (\ref{eq:tildeZoutformula1}) can be written as
3713: \be
3714: {\tilde J}_{mkn}(\lambda,\lambda),
3715: \ee
3716: where
3717: \begin{eqnarray}
3718: \fl
3719: {\tilde J}_{mkn}(\lambda_r,\lambda_\theta)
3720: = 2
3721: e^{-i m \phi_0} e^{i \omega_{mkn} t_0}
3722: \, e^{i(k \Upsilon_\theta \lambda_\theta + n \Upsilon_r
3723: \lambda_r)}
3724: \Theta_{\omega_{mkn}lm}[\theta(\lambda_\theta)]^*
3725: \nn \\ \mbox{} \times
3726: e^{- i m \Delta \phi_r(\lambda_r)} e^{-i m \Delta \phi_\theta(\lambda_\theta)}
3727: e^{i \omega_{mkn} \Delta t_r(\lambda_r)} e^{i
3728: \omega_{mkn} \Delta t_\theta(\lambda_\theta)}
3729: \nn \\ \mbox{}
3730: \times \left\{ G_{mkn}[\lambda_r,\lambda_\theta]
3731: R^{\rm out,down}_{\omega_{mkn}lm}[r(\lambda_r)]^*
3732: + G[\lambda_r,\lambda_\theta]
3733: \frac{d R^{\rm out,down}_{\omega_{mkn}lm}}{dr}[r(\lambda_r)]^*
3734: \right\}.
3735: \nonumber \\
3736: \label{eq:tildeJformula}
3737: \end{eqnarray}
3738: This function is biperiodic:
3739: \begin{eqnarray}
3740: {\tilde J}_{mkn}(\lambda_r + \Lambda_r, \lambda_\theta) &=&
3741: {\tilde J}_{mkn}(\lambda_r, \lambda_\theta) \nn \\
3742: \mbox{}
3743: {\tilde J}_{mkn}(\lambda_r , \lambda_\theta + \Lambda_\theta) &=&
3744: {\tilde J}_{mkn}(\lambda_r, \lambda_\theta).
3745: \end{eqnarray}
3746: Hence it can be expanded as a double Fourier series, and the average
3747: over $\lambda$ of ${\tilde J}_{mkn}(\lambda,\lambda)$ is
3748: \be
3749: \left< {\tilde J}_{mkn}(\lambda,\lambda) \right>_\lambda
3750: =\frac{1}{\Lambda_r \Lambda_\theta} \int_0^{\Lambda_r} d\lambda_r
3751: \int_0^{\Lambda_\theta} d\lambda_\theta {\tilde
3752: J}_{mkn}(\lambda_r,\lambda_\theta).
3753: \label{eq:Jav}
3754: \ee
3755: Substituting Eq.\ (\ref{eq:tildeJformula}) into Eq.\ (\ref{eq:Jav})
3756: and back into Eq.\ (\ref{eq:tildeZoutformula1}) now gives the formula
3757: (\ref{eq:tildeZlmknformula}).
3758:
3759:
3760: %\section{Evolution of orbits and asymptotic scalar waveform}
3761:
3762: \section{Circular orbits}
3763: \label{sec:circular}
3764:
3765:
3766: \subsection{Known formula for time derivative of the Carter constant}
3767:
3768:
3769: In this section we study the prediction of the formula (\ref{eq:dKdt})
3770: for the time derivative of the Carter constant, for the special case
3771: of circular orbits (orbits of constant Boyer-Lindquist radius $r$). It is known that
3772: circular orbits remain circular while evolving under the influence of radiation
3773: reaction in the adiabatic regime \cite{Kennefick:1995za,Ryan:1995xi,Thesis:Mino}.
3774: Also for circular orbits it is possible to relate the Carter constant
3775: $K$ to the energy $E$ and angular momentum $L_z$ of the orbit:
3776: specializing Eq.\ (\ref{eq:Kformula1}) to zero radial velocity $u_r
3777: =0$ yields
3778: \begin{equation}
3779: K = \frac{\left(\varpi^2 E - aL_z\right)^2}{\Delta} - r^2\;.
3780: \end{equation}
3781: In order for this relation to be preserved under slow evolution of
3782: $K$, $E$ and $L_z$, we must have
3783: \be
3784: \left< \frac{dK}{dt}\right>_t =
3785: \frac{2(\varpi^2 E - a L_z)}{\Delta}\left[\varpi^2 \left<
3786: \frac{dE}{dt} \right>_t - a
3787: \left< \frac{d L_z}{dt} \right>_t \right]\;.
3788: \label{eq:circtocirccheck}
3789: \ee
3790: We now verify that our expression (\ref{eq:dKdt}) for $\left< dK/dt
3791: \right>_t$ satisfies the condition (\ref{eq:circtocirccheck}).
3792:
3793:
3794: \subsection{Verification that our result reproduces this formula}
3795:
3796:
3797: For circular orbits, the following simplifications apply to the
3798: harmonic decomposition of the geodesic source derived in Sec.\
3799: \ref{sec:harmonic}. First, the functions $\Delta t_r(\lambda)$,
3800: ${\hat t}_r(\lambda)$, $\Delta \phi_r(\lambda)$ and ${\hat
3801: \phi}_r(\lambda)$ defined by Eqs.\ (\ref{eq:deltatrformula}),
3802: (\ref{eq:hattrdef}), (\ref{eq:deltaphirformula}) and
3803: (\ref{eq:hatphirdef}) vanish identically. Second,
3804: the biperiodic function $J_{\omega lm}(\lambda_r,\lambda_\theta)$ defined in Eq.\
3805: (\ref{eq:Jomegalmdef}) is therefore independent of $\lambda_r$. It follows that
3806: the double Fourier series (\ref{eq:doubleFourierseries})
3807: is replaced by the single Fourier series
3808: \be
3809: J_{\omega lm}(\lambda_\theta) =
3810: \sum_{k=-\infty}^\infty
3811: J_{\omega lmk} e^{-i k \Upsilon_\theta \lambda_\theta},
3812: \label{eq:singleFourierseries}
3813: \ee
3814: where the coefficients $J_{\omega lmk}$
3815: are given by
3816: \be
3817: J_{\omega lmk} = \frac{1}{\Lambda_\theta}
3818: \int_0^{\Lambda_\theta} d\lambda_\theta
3819: \,
3820: e^{i k \Upsilon_\theta \lambda_\theta}
3821: J_{\omega lm}(\lambda_\theta).
3822: \label{eq:Jomegalmkdef}
3823: \ee
3824: The effect of this change on the subsequent formulae for fluxes and
3825: amplitudes in Secs.\ \ref{sec:Edot} and \ref{sec:Kdot}
3826: can be summarized as: (i) remove the sums over $n$; (ii)
3827: remove the indices $n$ from the amplitudes $Z_{lmkn}$; (iii) remove
3828: the averaging operation
3829: \[
3830: \frac{1}{\lambda_r} \int_0^{\Lambda_r} d\lambda_r
3831: \]
3832: from the formulae for the amplitudes; and (iv) evaluate the
3833: integrands in the formulae for the amplitudes at $n=0$.
3834: Also the formula (\ref{eq:omegamkndef})
3835: for the frequency $\omega_{mkn}$ is replaced by
3836: \be
3837: \omega_{mk} = m \Omega_\phi + k \Omega_\theta.
3838: \ee
3839:
3840: With these simplifications, the formula (\ref{eq:dcalEdt2}) for the time
3841: derivatives of energy and angular momentum becomes
3842: \begin{eqnarray}
3843: \fl
3844: \left< \frac{d {\cal E}}{dt} \right>_t = - \frac{1}{64 \pi^2 \mu}
3845: \sum_{lmk}
3846: \left\{ \begin{array}{l} \omega_{mk} \\
3847: m \\
3848: \end{array} \right\}
3849: \frac{\omega_{mk}}{|\omega_{mk}|}
3850: \bigg[
3851: \frac{1}{|\alpha_{lmk}|^2}
3852: |Z^{\rm out}_{lmk} |^2
3853: \nonumber \\
3854: \lo
3855: +\frac{\omega_{mk} p_{mk}}{|\omega_{mk} p_{mk}|}
3856: \frac{|\tau_{lmk}|^2}{|\beta_{lmk}|^2}
3857: |Z^{\rm down}_{lmk} |^2 \bigg],
3858: \label{eq:dcalEdt2circ}
3859: \end{eqnarray}
3860: where ${\cal E} = E$ or $L_z$, and the expression in curly brackets
3861: means either $\omega_{mk}$ (for $E$) or $m$ (for $L_z$). Similarly
3862: the expression (\ref{eq:dKdt}) for the time derivative of the Carter constant
3863: becomes
3864: \begin{eqnarray}
3865: \fl
3866: \left< \frac{dK}{dt} \right>_t = - \frac{1}{32 \pi^2 \mu} \sum_{lmk}
3867: \frac{\omega_{mk}}{|\omega_{mk}|}
3868: \bigg[
3869: \frac{1}{|\alpha_{lmk}|^2}
3870: ( {\tilde Z}^{\rm out}_{lmk})^*
3871: Z^{\rm out}_{lmk}
3872: \nonumber \\
3873: \lo
3874: +\frac{\omega_{mk} p_{mk}}{|\omega_{mk} p_{mk}|}
3875: \frac{|\tau_{lmk}|^2}{|\beta_{lmk}|^2}
3876: ({\tilde Z}^{\rm down}_{lmk})^* Z^{\rm down}_{lmk}
3877: \bigg].
3878: \label{eq:dKdtcirc}
3879: \end{eqnarray}
3880: Comparing Eqs.\ (\ref{eq:dcalEdt2circ}) and (\ref{eq:dKdtcirc}), we
3881: see that the condition (\ref{eq:circtocirccheck})
3882: will be satisfied as long as the amplitudes ${\tilde Z}_{lmk}$ satisfy
3883: \be
3884: {\tilde Z}_{lmk}^{{\rm out},{\rm down}} =
3885: \frac{\left(\varpi^2 E- a
3886: L_z\right)}{\Delta}\left(\varpi^2\omega_{mk} - m a\right)
3887: Z_{lmk}^{{\rm out},{\rm down}}\;.
3888: \label{eq:amplitudeidentity}
3889: \ee
3890: Therefore it suffices to derive the identity
3891: (\ref{eq:amplitudeidentity}). We will derive this identity for the
3892: ``out'' modes; the ``down'' derivation is identical.
3893:
3894:
3895:
3896: The particular form of the expressions for the amplitudes that are
3897: most useful here are Eqs.\ (\ref{eq:Zoutuseful}) and
3898: (\ref{eq:tildeZoutformula1}):
3899: \be
3900: Z^{\rm out}_{lmk} = - \frac{2 \pi q}{\Gamma} \left< \Sigma \left(
3901: \pi_{lmk}^{\rm out} \right)^* \right>_\lambda,
3902: \label{eq:Zoutuseful1}
3903: \ee
3904: and
3905: \be
3906: {\tilde Z}^{\rm out}_{lmk} = \frac{2 \pi q}{\Gamma} \left<
3907: G_{mk}(\lambda,\lambda) \left( \pi_{lmk}^{\rm out} \right)^*
3908: + G(\lambda,\lambda) \partial_r \left( \pi_{lmk}^{\rm out} \right)^*
3909: \right>_\lambda.
3910: \label{eq:tildeZoutformula2}
3911: \ee
3912: Here we have defined $\pi_{lmk}^{\rm out} = \pi_{\omega_{mk}lm}^{\rm
3913: out}$. It follows from Eq.\ (\ref{eq:Gdef2}) together with
3914: $u_r=0$ that the function $G(\lambda,\lambda)$ vanishes, so the
3915: formula for ${\tilde Z}_{lmk}^{\rm out}$ simplifies to
3916: \be
3917: {\tilde Z}^{\rm out}_{lmk} = \frac{2 \pi q}{\Gamma} \left<
3918: G_{mk}(\lambda,\lambda) \left( \pi_{lmk}^{\rm out} \right)^*
3919: \right>_\lambda.
3920: \label{eq:tildeZoutformula3}
3921: \ee
3922: We now substitute into Eq.\ (\ref{eq:tildeZoutformula3})
3923: the expression (\ref{eq:Gmnkdef2}) for $G_{mk}$. The second term in
3924: Eq.\ (\ref{eq:Gmnkdef2}) vanishes since $u_r=0$, and of all the
3925: factors in the first term, only $\Sigma$ depends on $\lambda$; the
3926: remaining factors are constants and can be pulled out of the average
3927: over $\lambda$. This yields
3928: \be
3929: {\tilde Z}_{lmk}^{\rm out} = - \frac{2 \pi q}{\Gamma} \frac{\left( \varpi^2
3930: E - a L_z \right)}{\Delta} \left( \varpi^2 \omega_{mk} - a m \right) \left<
3931: \Sigma \left( \pi_{lmk}^{\rm out} \right)^* \right>_\lambda.
3932: \ee
3933: Comparing this with the formula (\ref{eq:Zoutuseful1}) for $Z_{lmk}^{\rm out}$ now
3934: yields the identity (\ref{eq:amplitudeidentity}).
3935:
3936:
3937: \ack
3938: This research was supported in part by NSF grants PHY-0140209
3939: and PHY-0244424 and by NASA Grant NAGW-12906. We thank the anonymous
3940: referees for several helpful comments.
3941:
3942:
3943: \appendix
3944:
3945:
3946: \section{Accuracy of adiabatic waveforms}
3947: \label{sec:accuracy}
3948:
3949: In this appendix we estimate the accuracy of the adiabatic waveforms,
3950: expanding on the brief treatment given in Ref.\ \cite{dfh05}.
3951: Specifically we estimate the magnitude of
3952: the correction to the
3953: waveform's phase that corresponds to the term $\phi_2(t)$ in Eq.\
3954: (\ref{eq:emri_phase}).
3955: We can roughly estimate this term by using post-Newtonian expressions
3956: for the waveform in which terms corresponding to $\phi_2(t)$ are readily identified.
3957: While the post-Newtonian approximation is not strictly valid in the
3958: highly relativistic regime near the horizon of interest here, it
3959: suffices to give some indication of the accuracy of the adiabatic
3960: waveforms. In this appendix we specialize for simplicity to circular,
3961: equatorial orbits. We also specialize to gravitational radiation
3962: reaction, unlike the body of the paper which dealt with the scalar case.
3963: Our conclusion is that adiabatic waveforms will likely be sufficiently accurate
3964: for signal detection.
3965:
3966:
3967:
3968:
3969: The procedure we use is as follows. We focus attention on the phase
3970: $\Psi(f)$ of the Fourier transform of the dominant, $l=2$ piece of the
3971: gravitational waveform. Here $f$ is the gravitational wave frequency.
3972: Now a change to the phase function of the form
3973: $\Psi(f) \to \Psi(f) + \Psi_0 + 2 \pi f \Delta t$ is not observable;
3974: the constant $\Delta t$ corresponds to a change in the time of arrival
3975: of the signal. Therefore we focus on the observable quantity $d^2
3976: \Psi/df^2$, and we write this as
3977: \be
3978: \frac{d^2 \Psi}{d f^2}(f) = G(f,m_1,m_2).
3979: \label{eq:Gadef}
3980: \ee
3981: Here $G$ is a function, $m_1$ is the mass of the black hole, and $m_2$
3982: is the mass of the inspiralling compact object\footnote{We also
3983: define $M = m_1 + m_2$, the total mass, and $\mu = m_1 m_2/M$, the
3984: reduced mass. Throughout the body of the paper we referred to the mass
3985: of the inspiralling object as $\mu$ and the mass of the black hole as
3986: $M$; this is valid to leading order in $\mu/M$. In this
3987: appendix however we need to be more accurate, and hence we revert to
3988: using $m_1$ and $m_2$ for the two masses.}. The specific form of $G$ obtained from
3989: post-Newtonian theory is discussed below.
3990:
3991:
3992: We next expand the function $G$ as a power
3993: series in $m_2$, at fixed $f$ and $m_1$. The leading order term scales as
3994: $m_2^{-1}$, and we obtain
3995: \be
3996: G(f,m_1,m_2) = \frac{G_0(f,m_1)}{m_2} + G_1(f,m_1) + G_2(f,m_1) m_2 + \ldots
3997: \label{eq:Gexpand}
3998: \ee
3999: The first term in this expression corresponds to the term $\Phi_1(t)$
4000: in Eq.\ (\ref{eq:emri_phase}); it gives the leading-order, adiabatic
4001: waveform. The error incurred from using adiabatic waveforms is
4002: therefore
4003: \be
4004: \fl
4005: \Delta G(f,m_1,m_2) = G(f,m_1,m_2) - \frac{G_0(f,m_1)}{m_2} =
4006: G_1(f,m_1) + G_2(f,m_1) m_2 + \ldots
4007: \ee
4008: We want to estimate the effects of this phase error.
4009:
4010:
4011: It is useful to split the phase error into two terms:
4012: \be
4013: \Delta G(f,m_1,m_2) = \Delta G_1(f,m_1,m_2) + \Delta G_2(f,m_1,m_2),
4014: \label{eq:DeltaGsplit}
4015: \ee
4016: where
4017: \be
4018: \Delta G_1(f,m_1,m_2) = G(f,m_1,m_2) - \frac{G_0(f,m_1 + m_2)}{m_2}
4019: \label{eq:DeltaG1def}
4020: \ee
4021: and
4022: \be
4023: \Delta G_2(f,m_1,m_2) = \frac{G_0(f,m_1 + m_2)}{m_2} - \frac{G_0(f,m_1)}{m_2}.
4024: \ee
4025: The phase error $\Delta G_2$ corresponds to the error that would be
4026: caused by using an incorrect value of the mass of the black hole. The
4027: effect of this error on the data analysis would be to cause a
4028: systematic error in the inferred best fit value of the black hole
4029: mass. However, the fractional error would be of order $\mu/M \ll 1$.
4030: This error will not have any effect on the ability of adiabatic
4031: waveforms to detect signals when used as search templates.
4032: Therefore, we will neglect this error, and focus on the remaining
4033: error term $\Delta G_1$.
4034:
4035:
4036: We now compute explicitly the phase error $\Delta \Psi_1(f)$ that
4037: corresponds to the term $\Delta G_1$ in Eq.\ (\ref{eq:DeltaGsplit}).
4038: We use the post-3.5-Newtonian expression for $\Psi(f)$, which
4039: can be computed from the orbital energy $E(f)$ and gravitational wave
4040: luminosity ${\dot E}(f)$
4041: via the equation \cite{Poisson:1995vs}
4042: \be
4043: \frac{d^2 \Psi}{df^2} = - 2 \pi \frac{dE/df(f)}{{\dot E}(f)}.
4044: \ee
4045: For the energy $E(f)$ we use the expression given in Eq.\ (50) of Ref.\
4046: \cite{Blanchet:1999pm}, and for the luminosity ${\dot E}$ we use Eq.\
4047: (12.9) of Ref.\ \cite{Blanchet:2001aw} \footnote{With parameter values $\lambda
4048: = -1987/3080$ from Ref.\ \protect{\cite{Blanchet:2003gy}} and $\theta =
4049: -11831/9240$ from Refs. \protect{\cite{Blanchet:2004bb,Blanchet:2004ek}}.}.
4050: This gives
4051: \be
4052: \Psi(f) = 2 \pi f t_c + \phi_c + \frac{3 M}{128 \mu y^5} {\cal F}(f),
4053: \label{eq:Psi0}
4054: \ee
4055: where $t_c$ and $\phi_c$ are constants, $y = (\pi M f)^{1/3}$, and
4056: \begin{eqnarray}
4057: \fl
4058: {\cal F} = 1 + \frac{3715\,y^2}{756} - 16\,\pi \,y^3 + \frac{15293365\,y^4}{508032} +
4059: \frac{11583231236531\,y^6}{4694215680} - \frac{6848\,\gamma\,y^6}{21}
4060: \nonumber \\
4061: - \frac{640\,{\pi }^2\,y^6}{3}
4062: + \frac{77096675\,\pi \,y^7}{254016} + \frac{55\,y^2\,\mu }{9\,M} +
4063: \frac{27145\,y^4\,\mu }{504\,M}
4064: \nonumber \\
4065: - \frac{15737765635\,y^6\,\mu }{3048192\,M} +
4066: \frac{2255\,{\pi }^2\,y^6\,\mu }{12\,M} + \frac{1014115\,\pi
4067: \,y^7\,\mu }{3024\,M} + \frac{3085\,y^4\,{\mu }^2}{72\,M^2}
4068: \nonumber \\
4069: + \frac{76055\,y^6\,{\mu }^2}{1728\,M^2} -
4070: \frac{36865\,\pi \,y^7\,{\mu }^2}{378\,M^2} -
4071: \frac{127825\,y^6\,{\mu }^3}{1296\,M^3}
4072: + \frac{a y^3}{3} \left(113 \nu^2 + \frac{75 \mu}{M} \right)
4073: \nonumber \\
4074: - \frac{13696\,y^6\,\log (2)}{21} + \frac{38645\,\pi \,y^5\,\log (y)}{252} -
4075: \frac{6848\,y^6\,\log (y)}{21}
4076: \nonumber \\
4077: + \frac{5\,\pi \,y^5\,\mu \,\log (y)}{M}.
4078: \label{eq:fullphase}
4079: \end{eqnarray}
4080: Here $\gamma$ is Euler's constant, $a$ is the dimensionless spin parameter of the
4081: black hole, and $\nu = m_1/M = 1/2 + \sqrt{1/4 -\mu/M}$.
4082: We have also added to Eq.\ (\ref{eq:fullphase}) the spin-dependent
4083: terms up to post-2-Newtonian order, taken from Eq.\ (1) of Ref.\
4084: \cite{Buonanno:2002fy}, specialized to a non-spinning
4085: particle on a circular, equatorial orbit.
4086:
4087:
4088:
4089:
4090: We now insert the expression (\ref{eq:Psi0}) for $\Psi(f)$ into Eqs.\
4091: (\ref{eq:Gadef}) and (\ref{eq:DeltaG1def}), and integrate twice with
4092: respect to $f$ to obtain $\Delta \Psi_1(f)$. The result is
4093: \begin{eqnarray}
4094: \fl
4095: \Delta \Psi_1(f) =
4096: \frac{3}{128 y^5} \bigg[
4097: \frac{55\,y^2}{9} +
4098: \frac{27145\,y^4 }{504} - \frac{15737765635\,y^6 }{3048192} +
4099: \frac{2255\,{\pi }^2\,y^6 }{12} + \frac{1014115\,\pi
4100: \,y^7 }{3024}
4101: \nonumber \\
4102: +\frac{3085\,y^4\,{\mu }}{72\,M}
4103: + \frac{76055\,y^6\,{\mu }}{1728\,M} -
4104: \frac{36865\,\pi \,y^7\,{\mu }}{378\,M} -
4105: \frac{127825\,y^6\,{\mu }^2}{1296\,M^2}
4106: \nonumber \\
4107: + 5\,\pi \,y^5 \,\log (y)
4108: + \frac{a y^3}{3} (75 + 113 q)
4109: \bigg],
4110: \label{eq:DeltaPsians}
4111: \end{eqnarray}
4112: where $q = (\nu^2-1) M/\mu = -2 + O(\mu/M)$.
4113: Now as discussed above, a change to the phase function of the form
4114: $\Psi(f) \to \Psi(f) + \Psi_0 + 2 \pi f \Delta t$ is not observable.
4115: This freedom allows us to set $\Psi(f_0) = \Psi'(f_0)
4116: = 0$
4117: %for all templates
4118: where $f_0$ is a fixed frequency,
4119: which amounts to
4120: replacing our expression $\Delta \Psi_1(f)$ for the phase error with
4121: \be
4122: \delta \Psi_1(f) = \Delta \Psi_1(f) - \Delta\Psi_1(f_0) - \Delta
4123: \Psi_1'(f_0)(f-f_0).
4124: \ee
4125:
4126: We now evaluate the phase error $\delta \Psi_1(f)$ for
4127: some typical sources. First, we take $\mu = 1 M_\odot$, $M = 10^6
4128: M_\odot$, $a =0.999$.
4129: For this case the last year of inspiral extends from $f = 10^{-2}$ Hz
4130: to $\sim 3 \times 10^{-2}$ Hz \cite{Finn:2000sy}; the maximum
4131: value of $|\delta \Psi_1|$ is $0.19$ cycles if we
4132: take $f_0 = 0.018$ Hz.
4133: As a second example with a higher mass ratio, we take $\mu = 10
4134: M_\odot$, $M = 10^6 M_\odot$, $a = 0.999$.
4135: The last year of inspiral for this case extends
4136: from $f = 0.004$ Hz to $f = 0.03$ Hz \cite{Finn:2000sy};
4137: the maximum
4138: value of $|\delta \Psi_1|$ is $0.47$ cycles if we
4139: take $f_0 = 0.0138$ Hz (see Fig.\ \ref{fig:phase})\footnote{If the
4140: spin-dependent term in Eq.\ (\protect{\ref{eq:DeltaPsians}}) is
4141: omitted, the errors are about $\sim 50\%$ larger.}. For signals from
4142: intermediate mass black holes that may be detectable by LIGO, the phase errors are
4143: yet smaller.
4144:
4145: \begin{figure}
4146: \begin{center}
4147: \epsfig{file=phase.eps,angle=0,width=9cm}
4148: \caption{This plot shows the error in the phase of the Fourier
4149: transform of the waveform that results from using adiabatic
4150: waveforms, estimated using post-3.5-Newtonian waveforms,
4151: for a $10 M_\odot$ compact object spiralling into a $10^6
4152: M_\odot$ black hole with spin parameter $a = 0.999$. For this system
4153: the last year of inspiral extends from $f = 0.004$ Hz to $f = 0.03$ Hz
4154: \cite{Finn:2000sy}.}
4155: \label{fig:phase}
4156: \end{center}
4157: \end{figure}
4158:
4159:
4160:
4161: These phase errors are sufficiently large that adiabatic waveforms
4162: cannot be used as data-analysis templates; the template will not quite
4163: stay in phase with the signal over the entire $\sim M/\mu \sim 10^5$ cycles of inspiral.
4164: However, the requirements on detection templates are much less
4165: stringent: phase coherence is only needed for $\sim 3$ weeks,
4166: rather than a year \cite{Gair:2004iv}. In addition, in the matched
4167: filtering search process phase error
4168: will tend to be compensated for by small systematic errors in
4169: the best-fit mass parameters. Because the adiabatic waveforms are
4170: {\it almost} good enough for data-analysis templates (phase errors
4171: $\sim1$ cycle in the worst cases), adiabatic waveforms will likely
4172: be accurate enough for detection templates both for ground-based and
4173: space-based detectors.
4174:
4175:
4176:
4177:
4178:
4179:
4180:
4181:
4182:
4183:
4184: \section{Typos in Gal'tsov}
4185: \label{sec:typos}
4186:
4187: This appendix lists some of the typos in the paper by Gal'tsov
4188: \cite{Galtsov:1982}:
4189:
4190: \begin{itemize}
4191:
4192: \item The discussion of the range of summation of the indices $l$, $m$
4193: given after Eq. (2.5) is incorrect. It should be summation $l$,$m$
4194: with $l \ge |s|$ and $|m| \le l$ rather than $l \le |s|$ and $|m|
4195: \le 1$.
4196:
4197: \item In the definition of the variable ${}_s u$ given just before
4198: Eq.\ (2.7), the factor $(\tau^2 + a^2)^{-1/2}$ should be $(r^2 +
4199: a^2)^{-1/2}$.
4200:
4201: \item In the first of the four equations in Eqs.\ (2.9), the term
4202: $r^2 e^{-i \omega r^*}$ should be $r^s e^{-i \omega r^*}$.
4203:
4204: \item The right hand side of Eq. (2.25) should be divided by
4205: $-4\pi$, as already discussed in Sec.\ \ref{sec:retardedformula} above.
4206:
4207: \item In Eq. (2.30), the quantity ${\bf f}_\Lambda^{\rm out}$ should
4208: be replaced by its complex conjugate ${\bar {\bf f}}_\Lambda^{\rm
4209: out}$.
4210:
4211: \item In the first line of Eq.\ (3.2), the quantity
4212: $Z(\theta',\varphi')$ should be replaced by its complex conjugate
4213: $Z(\theta',\varphi')^*$.
4214:
4215: \item In the second of Eqs.\ (3.3), the right hand side should read
4216: \[
4217: \kappa_s \tau_s {\bar \tau}_s (w k/|w k|) ( {}_sv^{\rm up} + {\bar
4218: \sigma}_{-s} \, {}_sv^{\rm in})
4219: \]
4220: rather than
4221: \[
4222: \kappa_s \tau_s {\bar \tau}_s (w k/|w k|) ( {}_sv^{\rm up}) + {\bar
4223: \sigma}_{-s} \, {}_sv^{\rm in},
4224: \]
4225: i.e. the closing bracket is in the wrong place.
4226:
4227: \item In the second term on the first line of Eq.\ (3.6), the first
4228: factor of ${}_s {\bar v}^{\rm up}(r)$ should be omitted.
4229:
4230: \item In Eq. (3.9), the argument of the last factor should be $x'$
4231: rather than $x x'$.
4232:
4233: \item Gal'tsov's equation (4.13) for the rate of change of energy or angular
4234: momentum of the orbit is correct only for sources which are smooth
4235: functions of frequency. However, for bound
4236: geodesics, the inner products $(\pi,J)$ are sums of delta functions
4237: in frequency, so it does not make sense to square these inner products
4238: as Gal'tsov does. The correct version of this equation is
4239: our Eq.\ (\ref{eq:dcalEdt2}) above.
4240:
4241: \end{itemize}
4242:
4243: \section*{References}
4244:
4245: %\bibliographystyle{hplain1}
4246: %\bibliography{scalar}
4247:
4248: \begin{thebibliography}{10}
4249:
4250: \bibitem{2002ApJ...581..438M}
4251: M.~C. {Miller}.
4252: \newblock {Gravitational radiation from intermediate-mass black holes}.
4253: \newblock {\em \apj}, 581:438--450, December 2002.
4254:
4255: \bibitem{Miller:2003sc}
4256: M.~Coleman Miller and E.~J.~M Colbert.
4257: \newblock Intermediate-mass black holes.
4258: \newblock {\em Int. J. Mod. Phys.}, D13:1--64, 2004, astro-ph/0308402.
4259:
4260: \bibitem{Fiorito:2004qh}
4261: Ralph Fiorito and Lev Titarchuk.
4262: \newblock Is {M82} {X-1} really an intermediate-mass black hole? {X}-ray
4263: spectral and timing evidence.
4264: \newblock 2004, astro-ph/0409416.
4265:
4266: \bibitem{Gair:2004iv}
4267: Jonathan~R. Gair et~al.
4268: \newblock Event rate estimates for {LISA} extreme mass ratio capture sources.
4269: \newblock 2004, gr-qc/0405137.
4270:
4271: \bibitem{2003ApJ...583L..21F}
4272: M.~{Freitag}.
4273: \newblock Gravitational waves from stars orbiting the {S}agittarius {A*} black
4274: hole.
4275: \newblock {\em Ap. J. Lett.}, 583:L21--L24, January 2003.
4276:
4277: \bibitem{Cutler:2002me}
4278: Curt Cutler and Kip~S. Thorne.
4279: \newblock An overview of gravitational-wave sources.
4280: \newblock {\em in {\it Proceedings of General Relativity and Gravitation XVI},
4281: edited by N.T. Bishop and S.D. Maharaj (Singapore, World Scientific)}, 2002,
4282: gr-qc/0204090.
4283:
4284: \bibitem{Poisson:1996tc}
4285: Eric Poisson.
4286: \newblock Measuring black-hole parameters and testing general relativity using
4287: gravitational-wave data from space-based interferometers.
4288: \newblock {\em Phys. Rev.}, D54:5939--5953, 1996, gr-qc/9606024.
4289:
4290: \bibitem{Barack:2003fp}
4291: Leor Barack and Curt Cutler.
4292: \newblock {LISA} capture sources: Approximate waveforms, signal-to- noise
4293: ratios, and parameter estimation accuracy.
4294: \newblock {\em Phys. Rev.}, D69:082005, 2004, gr-qc/0310125.
4295:
4296: \bibitem{Hughes:2002ei}
4297: Scott~A. Hughes and Roger~D. Blandford.
4298: \newblock Black hole mass and spin coevolution by mergers.
4299: \newblock {\em Astrophys. J.}, 585:L101--L104, 2003, astro-ph/0208484.
4300:
4301: \bibitem{Ryan:1995wh}
4302: F.~D. Ryan.
4303: \newblock Gravitational waves from the inspiral of a compact object into a
4304: massive, axisymmetric body with arbitrary multipole moments.
4305: \newblock {\em Phys. Rev.}, D52:5707--5718, 1995.
4306:
4307: \bibitem{Ryan:1997hg}
4308: Fintan~D. Ryan.
4309: \newblock Accuracy of estimating the multipole moments of a massive body from
4310: the gravitational waves of a binary inspiral.
4311: \newblock {\em Phys. Rev.}, D56:1845--1855, 1997.
4312:
4313: \bibitem{Finn:2000sy}
4314: Lee~Samuel Finn and Kip~S. Thorne.
4315: \newblock Gravitational waves from a compact star in a circular, inspiral
4316: orbit, in the equatorial plane of a massive, spinning black hole, as observed
4317: by {LISA}.
4318: \newblock {\em Phys. Rev.}, D62:124021, 2000, gr-qc/0007074.
4319:
4320: \bibitem{Glampedakis:2002cb}
4321: Kostas Glampedakis, Scott~A. Hughes, and Daniel Kennefick.
4322: \newblock Approximating the inspiral of test bodies into {Kerr} black holes.
4323: \newblock {\em Phys. Rev.}, D66:064005, 2002, gr-qc/0205033.
4324:
4325: \bibitem{Cutler:1994pb}
4326: C.~Cutler, D.~Kennefick, and E.~Poisson.
4327: \newblock Gravitational radiation reaction for bound motion around a
4328: {Schwarzschild} black hole.
4329: \newblock {\em Phys. Rev.}, D50:3816--3835, 1994.
4330:
4331: \bibitem{Shibata:1994xk}
4332: M.~Shibata.
4333: \newblock Gravitational waves by compact stars orbiting around rotating
4334: supermassive black holes.
4335: \newblock {\em Phys. Rev.}, D50:6297--6311, 1994.
4336:
4337: \bibitem{Glampedakis:2002ya}
4338: Kostas Glampedakis and Daniel Kennefick.
4339: \newblock Zoom and whirl: Eccentric equatorial orbits around spinning black
4340: holes and their evolution under gravitational radiation reaction.
4341: \newblock {\em Phys. Rev.}, D66:044002, 2002, gr-qc/0203086.
4342:
4343: \bibitem{Hughes:1999bq}
4344: Scott~A. Hughes.
4345: \newblock The evolution of circular, non-equatorial orbits of {Kerr} black
4346: holes due to gravitational-wave emission.
4347: \newblock {\em Phys. Rev.}, D61:084004, 2000, gr-qc/9910091.
4348:
4349: \bibitem{Teukolsky:1972my}
4350: S.~A. Teukolsky.
4351: \newblock Rotating black holes --- separable wave equations for gravitational
4352: and electromagnetic perturbations.
4353: \newblock {\em Phys. Rev. Lett.}, 29:1114--1118, 1972.
4354:
4355: \bibitem{Teukolsky:1973ha}
4356: Saul~A. Teukolsky.
4357: \newblock Perturbations of a rotating black hole. 1. {Fundamental} equations
4358: for gravitational electromagnetic, and neutrino field perturbations.
4359: \newblock {\em Astrophys. J.}, 185:635--647, 1973.
4360:
4361: \bibitem{Sasaki:1981sx}
4362: Misao Sasaki and Takashi Nakamura.
4363: \newblock Gravitational radiation from a {Kerr} black hole. 1. {Formulation}
4364: and a method for numerical analysis.
4365: \newblock {\em Prog. Theor. Phys.}, 67:1788, 1982.
4366:
4367: \bibitem{Mino:1997nk}
4368: Yasushi Mino, Misao Sasaki, and Takahiro Tanaka.
4369: \newblock Gravitational radiation reaction to a particle motion.
4370: \newblock {\em Phys. Rev.}, D55:3457--3476, 1997, gr-qc/9606018.
4371:
4372: \bibitem{Quinn:1997am}
4373: Theodore~C. Quinn and Robert~M. Wald.
4374: \newblock An axiomatic approach to electromagnetic and gravitational radiation
4375: reaction of particles in curved spacetime.
4376: \newblock {\em Phys. Rev.}, D56:3381--3394, 1997, gr-qc/9610053.
4377:
4378: \bibitem{Poisson:2003nc}
4379: Eric Poisson.
4380: \newblock The motion of point particles in curved spacetime.
4381: \newblock {\em Living Rev. Relativity}, 7:6, 2004, gr-qc/0306052.
4382:
4383: \bibitem{Krivan:1997hc}
4384: William Krivan, Pablo Laguna, Philippos Papadopoulos, and Nils Andersson.
4385: \newblock Dynamics of perturbations of rotating black holes.
4386: \newblock {\em Phys. Rev.}, D56:3395--3404, 1997, gr-qc/9702048.
4387:
4388: \bibitem{Burko:2002bt}
4389: Lior~M. Burko and Gaurav Khanna.
4390: \newblock Radiative falloff in the background of rotating black hole.
4391: \newblock {\em Phys. Rev.}, D67:081502, 2003, gr-qc/0209107.
4392:
4393: \bibitem{Scheel:2003vs}
4394: Mark~A. Scheel et~al.
4395: \newblock {3D} simulations of linearized scalar fields in {Kerr} spacetime.
4396: \newblock {\em Phys. Rev.}, D69:104006, 2004, gr-qc/0305027.
4397:
4398: \bibitem{Martel:2003}
4399: Karl Martel.
4400: \newblock {\em Ph.D. Thesis}.
4401: \newblock University of Guelph, 2003.
4402:
4403: \bibitem{Lopez-Aleman:2003ik}
4404: Ramon Lopez-Aleman, Gaurav Khanna, and Jorge Pullin.
4405: \newblock Perturbative evolution of particle orbits around {Kerr} black holes:
4406: Time domain calculation.
4407: \newblock {\em Class. Quant. Grav.}, 20:3259--3268, 2003, gr-qc/0303054.
4408:
4409: \bibitem{Khanna:2003qv}
4410: Gaurav Khanna.
4411: \newblock Teukolsky evolution of particle orbits around {Kerr} black holes in
4412: the time domain: Elliptic and inclined orbits.
4413: \newblock {\em Phys. Rev.}, D69:024016, 2004, gr-qc/0309107.
4414:
4415: \bibitem{Pazos-Avalos:2004rp}
4416: Enrique Pazos-Avalos and Carlos~O. Lousto.
4417: \newblock Numerical integration of the {Teukolsky} equation in the time domain.
4418: \newblock 2004, gr-qc/0409065.
4419:
4420: \bibitem{Rosenthal:2005ju}
4421: Eran Rosenthal.
4422: \newblock Regularization of second-order scalar perturbation produced by a
4423: point-particle with a nonlinear coupling.
4424: \newblock 2005, gr-qc/0501046.
4425:
4426: \bibitem{dfh05}
4427: Steve Drasco, Scott~A. Hughes, Eanna~E. Flanagan, and Joel Franklin.
4428: \newblock Gravitational radiation reaction and inspiral waveforms in the
4429: adiabatic limit.
4430: \newblock {\em to appear in Phys. Rev. Lett.}, gr-qc/0504015.
4431:
4432: \bibitem{Hinderer}
4433: Eanna~E. Flanagan and Tanja Hinderer.
4434: \newblock {Kerr} inspirals via a two timescale expansion.
4435: \newblock {\em in preparation}.
4436:
4437: \bibitem{Mino:2003yg}
4438: Yasushi Mino.
4439: \newblock Perturbative approach to an orbital evolution around a supermassive
4440: black hole.
4441: \newblock {\em Phys. Rev.}, D67:084027, 2003, gr-qc/0302075.
4442:
4443: \bibitem{Dirac:1938nz}
4444: Paul A.~M. Dirac.
4445: \newblock Classical theory of radiating electrons.
4446: \newblock {\em Proc. Roy. Soc. Lond.}, A167:148--169, 1938.
4447:
4448: \bibitem{Galtsov:1982}
4449: D.V. Gal'tsov.
4450: \newblock Radiation reaction in the {Kerr} gravitational field.
4451: \newblock {\em J. Phys A: Math. Gen.}, 15:3737, 1982.
4452:
4453: \bibitem{2005PThPh.113..733M}
4454: Yasushi {Mino}.
4455: \newblock Self-force in radiation reaction formula ---adiabatic approximation
4456: of the metric perturbation and the orbit.
4457: \newblock {\em Progress of Theoretical Physics}, 113:733--761, April 2005.
4458:
4459: \bibitem{2005Minob}
4460: Yasushi Mino.
4461: \newblock From the self-force problem to the radiation reaction formula.
4462: \newblock 2005, gr-qc/0506002.
4463:
4464: \bibitem{2005Minoc}
4465: Yasushi Mino.
4466: \newblock Extreme mass ratio binary: radiation reaction and gravitational
4467: waveform.
4468: \newblock 2005, gr-qc/0506008.
4469:
4470: \bibitem{tensorcase}
4471: Steve Drasco, Eanna~E. Flanagan, and Scott~A. Hughes.
4472: \newblock {Computing inspirals in Kerr in the adiabatic regime. II. The tensor
4473: case.}
4474: \newblock {\em in preparation}.
4475:
4476: \bibitem{Quinn:2000wa}
4477: Theodore~C. Quinn.
4478: \newblock Axiomatic approach to radiation reaction of scalar point particles in
4479: curved spacetime.
4480: \newblock {\em Phys. Rev.}, D62:064029, 2000, gr-qc/0005030.
4481:
4482: \bibitem{Drasco:2004tv}
4483: Steve Drasco and Scott~A. Hughes.
4484: \newblock Rotating black hole orbit functionals in the frequency domain.
4485: \newblock {\em Phys. Rev.}, D69:044015, 2004.
4486:
4487: \bibitem{Chrzanowski:1975wv}
4488: P.~L. Chrzanowski.
4489: \newblock Vector potential and metric perturbations of a rotating black hole.
4490: \newblock {\em Phys. Rev.}, D11:2042--2062, 1975.
4491:
4492: \bibitem{Gralla:2005et}
4493: Samuel~E. Gralla, John~L. Friedman, and Alan~G. Wiseman.
4494: \newblock Numerical radiation reaction for a scalar charge in {Kerr} circular
4495: orbit.
4496: \newblock 2005, gr-qc/0502123.
4497:
4498: \bibitem{Burko:2002ge}
4499: Lior~M. Burko, Abraham~I. Harte, and Eric Poisson.
4500: \newblock Mass loss by a scalar charge in an expanding universe.
4501: \newblock {\em Phys. Rev.}, D65:124006, 2002, gr-qc/0201020.
4502:
4503: \bibitem{Burko:2002gf}
4504: Lior~M. Burko.
4505: \newblock Instability of scalar charges in (1+1)d and (2+1)d.
4506: \newblock {\em Class. Quant. Grav.}, 19:3745--3752, 2002, gr-qc/0201021.
4507:
4508: \bibitem{Haas:2004kw}
4509: Roland Haas and Eric Poisson.
4510: \newblock Mass change and motion of a scalar charge in cosmological spacetimes.
4511: \newblock 2004, gr-qc/0411108.
4512:
4513: \bibitem{Wilkins:1972rs}
4514: Daniel~C. Wilkins.
4515: \newblock Bound geodesics in the {Kerr} metric.
4516: \newblock {\em Phys. Rev.}, D5:814--822, 1972.
4517:
4518: \bibitem{Detweiler:2002mi}
4519: Steven Detweiler and Bernard~F. Whiting.
4520: \newblock Self-force via a {Green's} function decomposition.
4521: \newblock {\em Phys. Rev.}, D67:024025, 2003, gr-qc/0202086.
4522:
4523: \bibitem{Brill:1972xj}
4524: D.~R. Brill, P.~L. Chrzanowski, C.~Martin~Pereira, E.~D. Fackerell, and J.~R.
4525: Ipser.
4526: \newblock Solution of the scalar wave equation in a {Kerr} background by
4527: separation of variables.
4528: \newblock {\em Phys. Rev.}, D5:1913--1915, 1972.
4529:
4530: \bibitem{drasconotes}
4531: Steve Drasco and Scott~A. Hughes.
4532: \newblock Gravitational wave snapshots of generic extreme mass ratio inspirals.
4533: \newblock {\em in preparation}.
4534:
4535: \bibitem{Kennefick:1995za}
4536: Daniel Kennefick and Amos Ori.
4537: \newblock Radiation-reaction-induced evolution of circular orbits of particles
4538: around {Kerr} black holes.
4539: \newblock {\em Phys. Rev.}, D53:4319--4326, 1996, gr-qc/9512018.
4540:
4541: \bibitem{Ryan:1995xi}
4542: Fintan~D. Ryan.
4543: \newblock Effect of gravitational radiation reaction on nonequatorial orbits
4544: around a {Kerr} black hole.
4545: \newblock {\em Phys. Rev.}, D53:3064--3069, 1996, gr-qc/9511062.
4546:
4547: \bibitem{Thesis:Mino}
4548: Yasushi Mino.
4549: \newblock PhD thesis, Kyoto University, 1996.
4550:
4551: \bibitem{Poisson:1995vs}
4552: Eric Poisson.
4553: \newblock Gravitational radiation from a particle in circular orbit around a
4554: black hole. 6. accuracy of the {post-Newtonian} expansion.
4555: \newblock {\em Phys. Rev.}, D52:5719--5723, 1995, gr-qc/9505030.
4556:
4557: \bibitem{Blanchet:1999pm}
4558: Luc Blanchet.
4559: \newblock Equations of motion of compact binaries at the third post-
4560: {Newtonian} order.
4561: \newblock {\em Pramana}, 53:1--15, 1999, gr-qc/0403122.
4562:
4563: \bibitem{Blanchet:2001aw}
4564: Luc Blanchet, Bala~R. Iyer, and Benoit Joguet.
4565: \newblock Gravitational waves from inspiralling compact binaries: Energy flux
4566: to third {post-Newtonian} order.
4567: \newblock {\em Phys. Rev.}, D65:064005, 2002, gr-qc/0105098.
4568:
4569: \bibitem{Blanchet:2003gy}
4570: Luc Blanchet, Thibault Damour, and Gilles Esposito-Farese.
4571: \newblock Dimensional regularization of the third {post-Newtonian} dynamics of
4572: point particles in harmonic coordinates.
4573: \newblock {\em Phys. Rev.}, D69:124007, 2004, gr-qc/0311052.
4574:
4575: \bibitem{Blanchet:2004bb}
4576: Luc Blanchet and Bala~R. Iyer.
4577: \newblock Hadamard regularization of the third {post-Newtonian} gravitational
4578: wave generation of two point masses.
4579: \newblock {\em Phys. Rev.}, D71:024004, 2005, gr-qc/0409094.
4580:
4581: \bibitem{Blanchet:2004ek}
4582: Luc Blanchet, Thibault Damour, Gilles Esposito-Farese, and Bala~R. Iyer.
4583: \newblock Gravitational radiation from inspiralling compact binaries completed
4584: at the third {post-Newtonian} order.
4585: \newblock {\em Phys. Rev. Lett.}, 93:091101, 2004, gr-qc/0406012.
4586:
4587: \bibitem{Buonanno:2002fy}
4588: Alessandra Buonanno, Yan-bei Chen, and Michele Vallisneri.
4589: \newblock Detecting gravitational waves from precessing binaries of spinning
4590: compact objects: Adiabatic limit.
4591: \newblock {\em Phys. Rev.}, D67:104025, 2003, gr-qc/0211087.
4592:
4593: \end{thebibliography}
4594:
4595:
4596:
4597: \end{document}
4598:
4599: