1: % what format do we use
2: \documentclass[aps,prd,showpacs,twocolumn,amssymb,eqsecnum,nofootinbib,floatfix]{revtex4}
3:
4: % packages
5: \usepackage{bm}
6: \usepackage{amsmath}
7: \usepackage{euscript}
8: \usepackage{graphicx}
9:
10: % use color package for editing information
11: \usepackage[usenames]{color}
12:
13: % make your own operator
14: \DeclareMathOperator{\sgn}{sgn}
15:
16: % enter the main document
17: \begin{document}
18:
19: \title{Verifying black hole orbits with gravitational spectroscopy}
20:
21: \author{Steve Drasco}
22: \affiliation{Jet Propulsion Laboratory, California Institute of
23: Technology, Pasadena, CA 91109}
24:
25: \date{\today}
26:
27: \begin{abstract}
28: Gravitational waves from test masses bound to geodesic orbits of rotating
29: black holes are simulated, using Teukolsky's black hole perturbation formalism,
30: for about ten thousand generic orbital configurations.
31: Each binary radiates power exclusively in modes with frequencies that are
32: integer-linear-combinations of the orbit's three fundamental frequencies.
33: General spectral properties are found with a survey of orbits about
34: a black hole taken to be rotating at 80\% of the maximal spin. The orbital eccentricity is varied
35: from 0.1 to 0.9. Inclination ranges from $20^\circ$ to $160^\circ$, and comes
36: to within $20^\circ$ of polar. Semilatus rectum is varied from 1.2 to 3 times the
37: value at the innermost stable circular orbits. The following general spectral properties are found:
38: (i) 99\% of the radiated power is typically carried by a few hundred modes, and at most by about a thousand modes,
39: (ii) the dominant frequencies can be grouped into a small number of families defined by
40: fixing two of the three integer frequency multipliers, and (iii) the specifics of these trends
41: can be qualitatively inferred from the geometry of the orbit under consideration.
42: Detections using triperiodic analytic templates modeled on these general properties
43: would constitute a verification of radiation from an adiabatic sequence of black hole orbits
44: and would recover the evolution of the fundamental orbital frequencies.
45: In an analogy with ordinary spectroscopy, this would compare to observing the Bohr model's atomic
46: hydrogen spectrum without being able to rule out alternative atomic theories or nuclei.
47: The suitability of such a detection technique is demonstrated using snapshots computed at
48: 12-hour intervals throughout the last three years before merger of a kludged inspiral.
49: The system chosen is typical of those thought to occur in galactic nuclei, and to be
50: observable with space-based gravitational wave detectors like LISA.
51: Because of circularization, the number of excited modes decreases as the binary evolves.
52: A hypothetical detection algorithm that tracks mode families dominating
53: the first 12 hours of the inspiral would capture 98\% of the total power over the remaining
54: three years.
55: \end{abstract}
56:
57: \pacs{04.70.-s, 97.60.Lf}
58:
59: \maketitle
60:
61: \section{Introduction} \label{Introduction}
62:
63:
64: The birth of physics in its modern form can arguably be placed at the
65: first successful efforts to monitor the motion of the planets as
66: they orbit the sun, and to model their planar elliptic orbits.
67: Since then, gravitation has transitioned from the best to the least
68: tested among the fundamental laws of physics. Gravitational spectroscopy,
69: monitoring power spectra of the outputs from gravitational wave detectors, can
70: verify the existence of two-body systems with a dramatically different character.
71: Athough characterized by mass ratios similar to those for planetary systems,
72: these systems are so distorted by the strong gravity near black hole event horizons that their orbits
73: look more like balls of twine than planar ellipses. These orbits differ significantly from
74: those with relativistic precessions which have already been observed and which may
75: soon be measured in our galactic center \cite{will 2007}. Given the relative
76: paucity of gravitational experiments, and the enigmatic status of gravitation compared to
77: the other fundamental interactions, opportunities to observe these systems are
78: highly valuable. Though unlikely candidates for Brahe-Kepler-like catalysts of another
79: revolution in physics, such observations would help to
80: elevate tests of gravity to a level that better compares with those for
81: the other fundamental physical interactions.
82:
83: General relativity predicts that black holes will radiate gravitational
84: waves characterized by discrete frequency spectra in at least two broad
85: classes of realistically observable scenarios: black hole
86: ringdowns, and during extreme mass ratio inspirals (EMRIs). In both, the
87: radiation is a consequence of a single black hole having been slightly perturbed.
88: Observations of this radiation are therefore ideally suited for studying
89: phenomena governed completely by the physics of isolated black holes.
90:
91: In the case of black hole ringdown, the source of the black hole's perturbation
92: is an arbitrary transient event, and the radiation is emitted as the hole
93: settles back into a quiescent state. Ringdown radiation consists of a
94: superposition of exponentially damped monochromatic waves, with frequencies and
95: damping timescales that are determined by a pair of continuous
96: parameters, the black hole's mass and spin. Potentially observable ringdown
97: events include binary systems that merge to form single black holes, and
98: supernovae that result in black hole formation. Ringdown phenomena have been
99: studied in detail elsewhere
100: \cite{flanagan hughes 1998, dreyer et al, berti et al 2006, berti et al 2007 ringdown},
101: and are not the focus of this work.
102:
103: This work instead focuses on EMRIs---black hole binary systems in which the
104: mass of the black hole is much greater than the mass of the companion.
105: The planned space-based interferometer LISA is anticipated to observe anywhere
106: from tens to thousands of EMRIs produced by the capture of compact
107: stellar mass black holes (and the occasional white dwarf) by the megamassive
108: black holes found in galactic nuclei (masses ranging from $10^5~M_\odot$ to $10^7~M_\odot$)
109: with sufficient sensitivity to determine the mass, spin, and quadrupole moment
110: of the larger black hole to within a fraction of a percent
111: \cite{gair et al 2004, hopman alexander 2006, miller et al 2005,
112: sigurdsson 2003, barack cutler 2004, barack cutler 2007}.
113: In the recent mock LISA data challenge, a proof of principal
114: for a variety of detection algorithms using simulated signals and noise,
115: three groups using independent detection algorithms recovered
116: EMRI masses and spins to within a few to a few tenths of a percent
117: \cite{gair mandel wen, gair babak porter barack, mldc}.
118:
119: Advanced ground-based detectors (with a lower frequency cutoff of about 10 Hz \cite{advanced ligo})
120: might also observe similar capture events by intermediate black holes in
121: globular clusters \cite{brown et al 2006}.
122: Observation of these intermediate mass ratio inspirals (IMRIs)
123: are more speculative, might occur with an estimated upper limit of a
124: few to a few tens per year \cite{mandel et al 2007}, and due to their
125: less extreme mass ratios, may require theoretical waveform models
126: that are more sophisticated than those based on black
127: hole perturbation theory alone.
128: See Ref.~\cite{amaro-seoane et al} for a review of EMRI science with both
129: ground-based and space-based detectors.
130:
131: Unlike radiation from black-hole ringdown, for EMRIs the source of the black
132: hole perturbation is persistent.
133: On timescales that are short compared to the observable lifetime of the binary,
134: the companion behaves as a test mass orbiting on a geodesic of the background
135: spacetime. Such orbits are in general characterized by three fundamental
136: frequencies \cite{schmidt 2002, mino 2003}, and the corresponding radiation is tri periodic,
137: or rather, a superposition of modes that oscillate exclusively at frequencies that
138: are integer-linear combinations of the orbit's three fundamental
139: frequencies \cite{drasco hughes 2004}
140: \begin{subequations}\label{waveform model}
141: \begin{align}
142: h_{+} - i h_{\times} &=
143: \sum_{m kn }
144: h_{mkn} e^{-i \omega_{mkn}t}~, & \\
145: \omega_{mkn} &= m \omega_\phi + k \omega_\theta + n \omega_r ~, \label{frequencies}&
146: \end{align}
147: \end{subequations}
148: where I am using the convention that all unconstrained summation indices range from $-\infty$ to $\infty$.
149: Here $\omega_\phi$, $\omega_\theta$, and $\omega_r$, are the fundamental orbital
150: frequencies associated with motion in the (Boyer-Lindquist) coordinate shown as a
151: subscript, $h_+$ and $h_\times$ are the two independent components of the metric
152: perturbation measured by a distant observer, and $h_{mkn}$ are complex amplitudes
153: that depend on the observer's position, and on the parameters of the binary
154: (masses, spin, and orbit geometry).
155:
156: The radiation described by Eq.~(\ref{waveform model}) is only a snapshot of a complete
157: EMRI waveform. While work toward precision waveforms for non-test-mass motion throughout
158: the lifetime of EMRIs remains an active field (see the discussion of Capra waveforms in Ref.~\cite{drasco 2006}),
159: for short enough times these waveform snapshots are as exact as any envisioned.
160: For example, for the case of a mass $\mu$ on a circular orbit of initial radius $r$ about a nonrotating
161: hole with mass $M \gg \mu$, the normalized overlap of the radiation from geodesic motion alone and that from the true inspiraling motion will be greater than 95\% for times shorter than \cite{drasco 2006}
162: \begin{equation} \label{dephasing time}
163: t_\text{dephase} \approx (1 \text{ day}) \left(\frac{M}{10^6 M_\odot}\right) \left(\frac{r}{6M}\right)^{11/4} \left(\frac{M/\mu}{10^5}\right)^{1/2} ~,
164: %t = M\left[ \sqrt{\frac{5 \pi}{108}} \left( \frac{r}{M} \right)^{11/4} \right] \sqrt{\frac{M}{\mu}} ~,
165: \end{equation}
166: whereas the orbital period is
167: \begin{equation}
168: t_\text{orbit} = (8 \text{ minutes}) \left(\frac{M}{10^6 M_\odot}\right) \left(\frac{r}{6M}\right)^{3/2}~.
169: %t = 2\pi M \left( \frac{r}{M} \right)^{3/2}
170: \end{equation}
171: Understanding the radiation from simple test mass motion may prove a sufficient basis for useful
172: observations, however crude. This is the motivating principal behind the work described here.
173:
174: This paper has two main results. The first is an observation of trends in
175: a survey of numerically simulated EMRI-snapshots, for thousands of different
176: generic orbital configurations, using the code described in
177: Refs.~\cite{hughes et al 2005} and \cite{drasco hughes 2006}. The observed
178: trends are summarized as follows:
179: (i) When the modes are sorted in order of decreasing power, their power
180: decreases fast enough to be primarily confined to a relatively small number of
181: modes compared to the number of modes computed when using the algorithm introduced in
182: Refs.~\cite{hughes et al 2005} and \cite{drasco hughes 2006}. (ii) The dominant
183: modes can be grouped into a small number of families defined by fixing two of the integer
184: frequency multipliers. (iii) The specifics of these trends (for
185: example, the precise falloff of power as a function of mode index, or the identity
186: of the dominant mode families) can be qualitatively inferred from the geometry
187: of the orbit under consideration, as illustrated by the following two examples.
188:
189: Simple orbits that are very nearly circular are expected to be typical of IMRIs
190: that could be found with ground-based detectors. Mandel \emph{et al}.~estimate
191: that the most likely IMRI formation mechanism should result in
192: an eccentricity $e < 10^{-4}$ by the time the orbital frequency is brought into
193: the observable band above 10 Hz for advanced LIGO \cite{advanced ligo}, and that even for the less likely
194: formations mechanisms that can yield higher eccentricities, 90\% of the systems
195: should have $e < 0.1$ at 10 Hz \cite{mandel et al 2007}.
196: For these orbits power falls off as a power law with mode index, and
197: 99\% of the radiated power is confined to a few to $\sim 10$ modes with frequencies
198: (\ref{frequencies}) in the following families:
199: \begin{equation}
200: \label{easy modes}
201: \omega = \left\{
202: \begin{array}{ll}
203: m\omega_\phi & \text{ dominant power at } m=2 \\
204: \pm m\omega_\phi + \omega_r & \text{ dominant power at } m=2 \\
205: \pm m\omega_\phi + \omega_\theta & \text{ dominant power at } m=1
206: \end{array}
207: \right. ~.
208: \end{equation}
209: where the upper (lower) sign of $\pm$ refers to prograde (retrograde) orbits, and
210: mode-power decreases exponentially as $m$ is varied away from
211: its value for the dominant mode. Note that for power spectra, there is no observational
212: consequence for a change in a frequency's overall sign. For any given spectrum, I
213: always define mode indices ($m,k,n$) in such a way that the frequency is positive: $\omega_{mkn} > 0$.
214:
215: More eccentric orbits are expected to be typical of the EMRIs
216: observed by space-based detectors. EMRIs that fall into LISA's frequency band are
217: thought to be born with such large initial eccentricities
218: ($10^{-6} \lesssim 1 - e \lesssim 10^{-3}$) that, although radiation circularizes them,
219: even at the time of merger as many as half of the systems should have residual eccentricities
220: of about $e \gtrsim 0.2$ \cite{barack cutler 2004}.
221: Increasing orbital eccentricity dramatically slows the rate of power falloff
222: and ultimately results in a spectrum dominated, albeit much less so, by
223: the following mode families:
224: \begin{equation}
225: \label{hard modes}
226: \omega = \left\{
227: \begin{array}{l}
228: \pm 2\omega_\phi + n\omega_r \\
229: \pm \omega_\phi + \omega_\theta + n\omega_r \\
230: \pm 2\omega_\theta + n\omega_r
231: \end{array}
232: \right. ~.
233: \end{equation}
234: where again $\pm$ is $+$ for prograde orbits and $-$ for retrograde orbits,
235: and mode power decreases as $n$ is varied away from its value $n_{\max}$ at which the
236: spectra are peaked. The value of $n_{\max}$ can be very crudely predicted, with an accuracy
237: on the order of 10\%, from eccentricity $e$ according to the formula
238: \begin{equation} \label{PM peaks}
239: n_{\max} \approx \exp(1/2) \left( 1-e\right)^{-3/2}~,
240: \end{equation}
241: that was fitted \cite{drasco hughes 2006} to the peaks in the spectra
242: derived by Peters and Mathews, who in 1963 used an analytical treatment of
243: Newtonian orbits with the quadrupole formula for gravitational radiation \cite{peters mathews}.
244: For these eccentric orbits, the rate of power falloff as a function of mode index is
245: such that as many as $\sim 10^2$ to $10^3$ modes are needed to capture 99\% of the power radiated.
246:
247: The second main result of the paper is a proposal for combining the analytical
248: waveform model (\ref{waveform model}) with the observed spectral trends
249: so as to create a means for verifying fundamental aspects of black hole physics.
250: Precision measurements that
251: will test the fundamental physics of black holes have long been among the primary
252: motivations for gravitational wave observatories \cite{abramovici et al, lisa science}.
253: However, whether or not there are any feasible means for implementing one of the
254: more grand measurements, measuring the multipolar
255: decomposition of spacetime near black hole candidates in such a way as to identify general relativity as the
256: only valid theory, has been a subject of
257: debate \cite{hughes 2006, psaltis et al 2007}.
258: In its most pure form, that measurement will require more than the current
259: understanding of what exactly is predicted by general relativity and its
260: alternatives. I propose a related, but less ambitious measurement that will
261: constitute a minimal verification of one of general relativity's fundamental predictions
262: for black hole perturbation: that black holes perturbed by bound test masses
263: radiate according to the above waveform model (\ref{waveform model}) for times
264: that are long compared to the spectrum's fundamental periods
265: $2\pi / \omega_{\phi, \theta, r}$ but short compared to the inspiral time. The measurement would also
266: recover the evolutionary sequence of those three fundamental frequencies.
267: This is a minimal verification in that it only provides a means for confirming general
268: relativity's prediction under the assumption that the perturbed spacetime
269: geometry is that of a rotating black hole.
270: Observations that provided such a verification may ultimately be subjected to
271: the more grand tests based on generalizations \cite{li lovelace 2007} of
272: Ryan's theorem \cite{ryan}. In the absence of a greater theoretical understanding
273: of alternatives to general relativity and black holes ,however, the verification alone
274: would not disqualify alternatives for either the background
275: spacetime or the theory of gravity.
276:
277: The remaining sections of this paper are outlined as follows. In
278: Sec.~\ref{EMRI-snapshot spectra}, I review the relevant equations that define
279: generic black hole orbits and that describe the radiation they produce. In
280: Sec.~\ref{Sample spectra}, the radiative spectra for three sample configurations
281: are studied and general trends are discussed.
282: In Sec.~\ref{Survey of many spectra} I simulate spectra from a grid of orbital configurations
283: and discuss how the general trends from the previous section vary across the grid.
284: Section \ref{Spectra from a kludged inspiral}
285: describes how the spectral trends evolve throughout the lifetime of an approximated inspiral.
286: In Sec.~\ref{Verification of black hole orbits},
287: I outline a practical means for extracting signals characterized by these
288: general spectral trends from data collected by gravitational wave observatories.
289: Section \ref{Conclusion} summarizes the paper's main points.
290:
291:
292: \section{EMRI-snapshot spectra}\label{EMRI-snapshot spectra}
293:
294: Radiation from generic configurations of test masses bound to black holes has
295: been studied in previous work \cite{drasco flanagan hughes 2005, drasco hughes 2006}.
296: In this section, I review definitions and equations derived there which will be needed
297: throughout the remaining sections.
298:
299: \subsection{Orbits}
300:
301: The physical system described by an EMRI waveform snapshot is a
302: nonspinning\footnote{For the EMRIs that could be seen with LISA, the mass ratio renders
303: the spin of the smaller object negligible (see Appendix C of Ref.~\cite{barack cutler 2004}).
304: The same may not be true for the IMRIs that could be seen with advanced ground-based
305: detectors, but the effect would likely be competing with more significant complications
306: due to the less extreme mass ratios for IMRIs.}
307: test mass $\mu$ bound to a black hole with mass $M$ and a spin per unit mass of
308: magnitude $0 \le a \le M$. The test mass' orbit is a bound geodesic of the Kerr
309: spacetime determined by $M$ and $a$. In Boyer-Lindquist coordinates
310: $(t, r, \theta, \phi)$, the orbit as a function of proper
311: time $x^\mu(\tau)$ satisfies the four first order geodesic
312: equations derived by Carter \cite{mtw, carter}. When bound solutions are
313: parametrized as functions of Mino time $\lambda$, with
314: $d\tau = (r^2 + a^2 \cos^2\theta)d\lambda$, two of
315: their coordinates are periodic \cite{mino 2003}
316: \begin{subequations} \label{mino orbits}
317: \begin{align}
318: t(\lambda) &= \Gamma\lambda
319: + \sum_{kn}
320: t_{kn} e^{-i(k \Upsilon_\theta + n \Upsilon_r) \lambda}~,& \\
321: r(\lambda) &=
322: \sum_{n}
323: r_{n} e^{-in \Upsilon_r \lambda}~,& \\
324: \theta(\lambda) &=
325: \sum_{k}
326: \theta_{k} e^{-ik \Upsilon_\theta \lambda}~,& \\
327: \phi(\lambda) &= \Upsilon_\phi \lambda
328: + \sum_{kn}
329: \phi_{kn} e^{-i(k \Upsilon_\theta + n \Upsilon_r)\lambda}~,&
330: \end{align}
331: \end{subequations}
332: where the coefficients in front of the exponentials are constants (with values that cause these seemingly complex sums to be real).
333: The quantity $\Gamma$ relates the Mino-frequencies $\Upsilon_{r,\theta,\phi}$ to
334: coordinate-time frequencies
335: \begin{equation}
336: \omega_{r,\theta, \phi} = \Upsilon_{r, \theta, \phi} / \Gamma ~,
337: \end{equation}
338: that appear in the radiation observed by distant observers (\ref{waveform model}). The three
339: spatial Boyer-Lindquist coordinates of the orbit
340: are not periodic functions of coordinate time $t$, however it follows
341: from the formalism in Ref.~\cite{drasco hughes 2004} that
342: they have simple biperiodic forms
343: \begin{subequations} \label{coordinate orbits}
344: \begin{align}
345: r(t) &=
346: \sum_{kn}
347: {\tilde r_{kn}} e^{-i(k \omega_\theta + n \omega_r)t}~,& \\
348: \theta(t) &=
349: \sum_{kn}
350: {\tilde \theta_{kn}} e^{-i(k \omega_\theta + n \omega_r)t}~,& \\
351: \phi(t) &= \omega_\phi t
352: + \sum_{kn}
353: {\tilde \phi_{kn}} e^{-i(k \omega_\theta + n \omega_r)t}~,&
354: \end{align}
355: \end{subequations}
356: where the expansion coefficients are again constants.
357: A derivation of the coefficients ${\tilde r_{kn}}$,
358: ${\tilde \theta_{kn}}$, and ${\tilde \phi_{kn}}$ in terms of
359: the coefficients in the Mino-time series (\ref{mino orbits}) is given in
360: Appendix \ref{append}.
361:
362: The orbital frequencies $\omega_{r,\theta,\phi}$ are uniquely determined by
363: specifying the three constants associated with Killing fields, energy $E$, axial
364: angular momentum $L$, and Carter's constant $Q$
365: \begin{align}
366: E &= - \mu u^t ~,& \\
367: L &= \mu M u^\phi ~,& \\
368: Q &= (r^2 + a^2\cos^2\theta)^2 (u^\theta)^2 & \nonumber \\
369: &+ L^2 \cot^2 \theta + a^2(\mu^2 - E^2) \cos^2 \theta~,&
370: \end{align}
371: where $u^\alpha = dx^\mu/d\tau$ is the orbit's four-velocity.
372: They can also be determined by
373: specifying the orbit's coordinate boundaries between two radial turning
374: points and between two angular turning points that are symmetric about
375: the equatorial plane at $\theta = \pi/2$
376: \begin{align}
377: r_{\min} &\leq r \leq r_{\max}~,& \\
378: \theta_{\min} &\leq \theta \leq \pi - \theta_{\min}~,&
379: \end{align}
380: or by specifying
381: three geometric constants generalized from Newtonian orbits:
382: eccentricity $e$, semilatus rectum $p$, and inclination $\iota$
383: \begin{align}
384: \iota & = \frac{\pi}{2} - \sgn(\omega_\phi) \theta_{\min}~,& \\
385: \frac{r_{\min}}{M} &= \frac{p}{1 + e}~,& \\
386: \frac{r_{\max}}{M} &= \frac{p}{1 - e}~,&
387: \end{align}
388: where $\sgn(\omega_\phi)$ is $1$ for prograde orbits and $-1$ for retrograde orbits.
389: See Appendix A of Refs.~\cite{schmidt 2002} or \cite{drasco hughes 2006} for
390: explicit formulae relating the geometric orbital constants,
391: the formal Killing constants $E$, $L$, and $Q$, the frequencies
392: $\omega_{\phi,\theta,r}$, and the Mino frequencies $\Gamma$ and $\Upsilon_{\phi, \theta, r}$.
393: Each of the following sets of parameters are uniquely determined by fixing the values
394: for any one of them
395: \begin{align}
396: &(\Upsilon_\phi, \Upsilon_\theta, \Upsilon_r)~, & \\
397: &(\omega_\phi, \omega_\theta, \omega_r)~,& \\
398: &(E, L, Q)~, & \\
399: &(r_{\min}, r_{\max}, \theta_{\min})~, & \\
400: &(e, p, \iota)~.&
401: \end{align}
402: Following the terminology of the Guelph group
403: \cite{pound poisson nickel 2005, pound poisson 2007a, pound poisson 2007b}, each of
404: these triples is a complete set of principal orbital elements.
405:
406: After fixing the principal orbital elements, the orbit is not completely determined
407: until one specifies an initial position, or some equivalent set of parameters, called
408: positional orbital elements \cite{pound poisson nickel 2005, pound poisson 2007a, pound poisson 2007b}.
409: Here I will use the following orbital elements
410: \begin{equation}
411: (\lambda_t, \lambda_\phi, \lambda_r, \lambda_\theta)~.
412: \end{equation}
413: These are defined such that, after specifying them and the principal orbital elements, any bound
414: black hole orbit can be uniquely expressed as follows
415: \begin{subequations} \label{general from fiducial}
416: \begin{align}
417: t(\lambda) &= \Gamma(\lambda - \lambda_t)
418: + \sum_{k=1}^\infty \hat t_k^\theta
419: \sin[k\Upsilon_\theta(\lambda-\lambda_\theta)]
420: \nonumber & \\
421: & + \sum_{n=1}^\infty \hat t_n^r
422: \sin[n\Upsilon_r(\lambda-\lambda_r)] ~,& \\
423: r(\lambda) &=
424: \sum_{n=0}^\infty
425: \hat r_{n} \cos[n \Upsilon_r (\lambda-\lambda_r)] ~,& \\
426: \theta(\lambda) &=
427: \sum_{k=0}^\infty
428: \hat \theta_{k} \cos[k \Upsilon_\theta (\lambda-\lambda_\theta)] ~,& \\
429: \phi(\lambda) &= \Upsilon_\phi (\lambda - \lambda_\phi)
430: + \sum_{k=1}^\infty \hat \phi_k^\theta
431: \sin[k\Upsilon_\theta(\lambda-\lambda_\theta)]
432: \nonumber & \\
433: & + \sum_{n=1}^\infty \hat \phi_n^r
434: \sin[n\Upsilon_r(\lambda-\lambda_r)] ~,&
435: \end{align}
436: \end{subequations}
437: were the hatted coefficients depend only on the principal orbital elements and are given by
438: integrals over the $\lambda$-derivatives of the coordinates as given by Eqs.~(2.27) and (2.28) in
439: Ref.~\cite{drasco hughes 2006}\footnote{Because of a difference in notation, replace
440: $\Delta x$ there with $\hat x$ used here, for $x = t, \phi$.}.
441: The first two positional elements, $\lambda_t$ and $\lambda_\phi$ can take on any value. The second two are defined on
442: $0 \leq \lambda_{r} < 2\pi/\Upsilon_r$ and $0 \leq \lambda_{\theta} < 2\pi/\Upsilon_\theta$, respectively as the
443: smallest positive values of $\lambda$ at which the coordinate shown in their subscript reaches the smaller of its two turning points.
444:
445: \subsection{Radiation}
446:
447: Black holes that are forever perturbed by bound test masses produce
448: a gravitational wave field whose two orthogonal linearly polarized components, $h_+$ and $h_\times$, can be expressed as a single complex function made up of a series of modes that oscillate at frequencies that are integer-linear combinations of the orbit's fundamental frequencies \cite{drasco hughes 2004}. For
449: an observer located at $(t, r, \theta, \phi)$, that function is
450: \begin{equation}
451: h_{+} - i h_{\times} = \sum_{mkn} h_{mkn} e^{-i \omega_{mkn}t}~,
452: \end{equation}
453: up to corrections of order $\mu^2/M^2$. The complex mode amplitudes are given by
454: \begin{equation} \label{mode amplitudes}
455: h_{mkn} = -2\frac{e^{i (\omega_{mkn}r + m\phi)}}{r\omega^2_{mkn}}
456: \sum_{l = 2}^\infty
457: Z_{lmkn}
458: S_{lmkn}(\theta)
459: \end{equation}
460: where $S_{lmkn}(\theta)$ and $Z_{lmkn}$ are quantities that are found by applying Teukolsky's black
461: hole perturbation formalism \cite{teukolsky}, as described in Sec.~III of Ref.~\cite{drasco hughes 2006}. The
462: real function $S_{lmkn}(\theta)$, that has here been normalized over the unit sphere
463: \begin{equation}
464: 2\pi \int_0^\pi d\theta~ [S_{lmkn}(\theta)]^2 \sin\theta = 1~,
465: \end{equation}
466: satisfies Teukolsky's angular equation, a homogeneous ordinary differential equation.
467: The complex numbers $Z_{lmkn}$ are constant coefficients that come from the limiting behavior
468: of the physical solutions to Teukolsky's radial equation, an inhomogeneous ordinary differential equation.
469: At infinity, that equation's general solutions have the following
470: form (see Teukolsky's original papers \cite{teukolsky}, or for a particularly good textbook-level
471: discussion, Sec.~4.8.1 of \cite{andersson frolov novikov})
472: \begin{equation}
473: R_{lmkn}(r\to \infty) = R^{\text{out}}_{lmkn}r^3 e^{i \omega_{mkn} r^*}
474: + R^{\text{in}}_{lmkn} \frac{1}{r} e^{-i\omega_{mkn} r^*}~.
475: \end{equation}
476: The solution of Teukolsky's master equation fully describes the system's radiation, to first order in the perturbation, and is a sum over products of the separated functions
477: $R_{lmkn}(r)$, $S_{lmkn}(\theta)$, $e^{im\phi}$, and $e^{-i\omega_{mkn}t}$.
478: Therefore, at radial infinity, the complex coefficients $R^\text{out}_{lmkn}$ determine the amount of outgoing radiation (toward infinity), while the $R^\text{in}_{lmkn}$ determine the amount of in-going radiation (toward the event horizon). The complex numbers $Z_{lmkn}$ appearing in the mode amplitudes for the gravitational
479: wave (\ref{mode amplitudes}) are taken from the solution that obeys a boundary condition with zero ingoing radiation at infinity
480: \begin{equation}
481: R_{lmkn}(r\to\infty) = Z_{lmkn} r^3 e^{i\omega_{mkn}r}~,
482: \end{equation}
483: as well as only in-going radiation at the event horizon. Details on how to calculate these solutions, and on how the
484: horizon-boundary condition enters, can be found in Ref.~\cite{drasco hughes 2006}.
485:
486: The time-averaged rates of change for the principal orbital elements can be given in terms of
487: the quantities described above\footnote{The time averages can be simply defined in terms of Mino-time
488: integrals of length $2\pi/\Upsilon_{r}$ and $2\pi/\Upsilon_{\theta}$. See Sec.~3.8 and 9.1 of Ref.~\cite{drasco flanagan hughes 2005} for details.}.
489: Each rate of change can be expressed as two fluxes: an outgoing
490: flux at radial infinity and an ingoing flux at the event horizon. Horizon fluxes are usually two or three orders of magnitude smaller than the fluxes at infinity (see Ref.~\cite{poisson sasaki} for an analytic treatment of circular nonspinning binaries, or Tables IV and VII from Ref.~\cite{drasco hughes 2006} for
491: numerical examples of generic binaries). Though they may prove
492: observable indirectly through their influence on the orbital evolution \cite{li lovelace 2007}, I will not
493: discuss them here apart from mentioning that their calculation is
494: nearly identical to that of fluxes at infinity.
495:
496: The time-averaged power radiated to infinity at frequency $\omega_{mkn}$, as measured by distant observers,
497: is given by \cite{drasco hughes 2006}
498: \begin{equation}
499: \left< \frac{dE}{dt} \right>_{mkn} = \frac{1}{4\pi \omega_{mkn}^2} \sum_{l = 2}^\infty
500: \left| Z_{lmkn} \right|^2 ~,
501: \end{equation}
502: and the total time-averaged energy flux at infinity is then just a sum of the power radiated
503: at all possible frequencies
504: \begin{equation}
505: \left< \frac{dE}{dt} \right> = \sum_{mkn} \left< \frac{dE}{dt} \right>_{mkn}~.
506: \end{equation}
507: The remainder of this paper studies the distribution of power among the various terms in this sum
508: and how that distribution is affected by the configuration of the orbit. Understanding this
509: distribution is akin to understanding the evolution of the other princple orbital elements, since
510: they are all somewhat simply related. The time-averaged flux of the other two formal constants of
511: motion can be written as \cite{drasco flanagan hughes 2005, sago et al 2005, drasco hughes 2006, sago et al 2006}
512: \begin{eqnarray} \label{fluxes}
513: \left< \frac{dL}{dt} \right> &=& \sum_{mkn} \left< \frac{dE}{dt} \right>_{mkn}
514: \frac{m}{\omega_{mkn}}\\
515: \left< \frac{dQ}{dt} \right> &=& 2\sum_{mkn} \left< \frac{dE}{dt} \right>_{mkn}
516: \left(
517: \frac{m}{\omega_{mkn}} L \left<\cot^2 \theta \right> \right. \nonumber \\
518: &+&\left. \frac{k}{\omega_{mkn}} \mu\Upsilon_\theta
519: -a^2 E \left<\cos^2 \theta \right>
520: \right)~,
521: %\left< \frac{dQ}{dt} \right> &=& 2\sum_{mkn} \left< \frac{dE}{dt} \right>_{mkn}
522: % \left(
523: % a^2 E \left<\cos^2 \theta \right> \right. \nonumber \\
524: % &+&\left. \frac{m}{\omega_{mkn}} L \left<\cot^2 \theta \right>
525: % +\frac{k}{\omega_{mkn}} \Upsilon_\theta
526: % \right)~,
527: \end{eqnarray}
528: where the angled brackets represent a time average.
529: From these expressions, one can compute the time-averaged evolution for any
530: set of principal orbital elements due to radiation at infinity.
531:
532: This all but concludes the review of quantities needed here to describe EMRI snapshots.
533: Before moving on however, an important property about how the various orbital parameters
534: influence the radiation should be addressed. It has been shown that the positional orbital
535: elements have a somewhat simple influence on the radiation, by comparison
536: to the influence of the principal orbital elements. Their only influence on the quantities
537: discussed above is as a phase factor in the complex numbers $Z_{lmkn}$. That factor can be written as \cite{drasco flanagan hughes 2005, mino 2007}
538: \begin{equation} \label{positional phase}
539: Z_{lmkn}(\lambda_t, \lambda_\phi, \lambda_{r}, \lambda_{\theta}) = e^{i \chi_{mkn}}
540: Z_{lmkn}(0, 0, 0, 0) ~,
541: \end{equation}
542: where $\chi_{mkn}$ is given by
543: \begin{equation}
544: \chi_{mkn} = m \Upsilon_\phi (\lambda_\phi - \lambda_t)
545: + k \Upsilon_\theta(\lambda_\theta - \lambda_t)
546: + n \Upsilon_r(\lambda_r - \lambda_t)~.
547: \end{equation}
548: The radiative fluxes for the principal orbital elements (\ref{fluxes}) are therefore independent
549: of the positional orbital elements, since they depend only on $|Z_{lmkn}|$. So the positional orbital elements
550: will not be relevant to the discussion of power spectra and their evolution in the remaining sections.
551: However, the mode amplitudes of the waveform (\ref{mode amplitudes}) are functions of
552: $Z_{lmkn}$, and not just their moduli. Therefore, knowing the evolution of the principle orbital
553: elements alone is insufficient for evolving from one snapshot to the next in an optimal coherent matched filtering
554: detection algorithm. To evolve waveforms coherently, one needs a prescription for changing both the principal and positional orbital elements. This issue will be revisited when discussing detection algorithms in Sec.~\ref{Verification of black hole orbits}.
555:
556: Before discussing the results of the simulations, it is perhaps useful to remind readers who are
557: more familiar with other simulations of radiating black hole binary systems that for these snapshot spectra, the source of the radiation is ever-present. There is no initial data from which imperfections could produce junk radiation which dies out over time. Of course no source of radiation could really persist forever like this, but that is why these spectra are ``snapshots'' of EMRI spectra. As described in the introduction, the snapshot spectra match
558: the true spectra from EMRIs only for sufficiently short observation times, shorter than the dephasing time
559: (\ref{dephasing time}).
560:
561: \section{Sample spectra}\label{Sample spectra}
562:
563: In this section I describe EMRI snapshot spectra from three sample systems and
564: identify properties that are useful for understanding spectra from systems with arbitrary
565: orbit geometries. All of the spectra shown in this paper
566: were simulated using the numerical code that was first described in Ref.~\cite{drasco hughes 2006}.
567: When simulating the spectra described in this paper,
568: all adjustable parameters of that code were set to the same values
569: used when computing the catalog of orbits introduced in Sec.~V of that paper, with one exception.
570: The one exceptional code parameter is the requested fractional accuracy $\varepsilon_\text{flux}$
571: in the total radiated power $\left<dE/dt\right>$. When relevant, the value of $\varepsilon_\text{flux}$ used here
572: will be stated below.
573:
574: Figure~\ref{simple spectra} shows the dominant spectral lines from two relatively simple orbits, both
575: computed to a fractional accuracy of $\varepsilon_\text{flux} = 10^{-6}$, in the total
576: radiated power $\left<dE/dt\right>$. The orbits for these two spectra are simple in the sense that
577: the motion of the test mass is very nearly restricted to a constant radius, and to the equatorial plane of the large hole. Correspondingly, these spectra are also somewhat simple.
578: \begin{figure*}
579: \includegraphics[width = .459\textwidth]{Spectrum_1.eps}
580: \includegraphics[width = .459\textwidth]{Spectrum_2.eps}
581: \caption{\label{simple spectra} The dominant spectral lines for two relatively simple orbital configurations.
582: The parameters describing the system (black hole spin $a$, eccentricity $e$, semilatus rectum
583: $p$, and inclination $\iota$) are shown above each panel. For the spectrum on the left, 80\% of the total power is
584: carried by the dominant mode and 99\% is carried by the most powerful 6 modes.
585: For the spectrum on the right, 77\% of the total power is
586: carried by the dominant mode, and 99\% is carried by the most powerful 11 modes.
587: For each spectrum, 99.99\% of the total power is carried by the lines shown.
588: The red solid line traces over the tops of the lines carrying radiation at frequencies
589: $\omega_{m00}= m\omega_\phi$, for various values of $m$ (which can be read off from the horizontal axis).
590: Similarly, the green dash-dot line
591: highlights frequencies $\omega_{m10}= m\omega_\phi+\omega_\theta$ (peaked at $m = 1$), and the
592: dashed blue line shows $\omega_{m01}= m\omega_\phi+\omega_r$ (peaked at $m=2$). The thicker spectral
593: lines are members of these three mode families, while the thinner lines are not.}
594: \end{figure*}
595: For both, the peak in the power spectrum occurs at a frequency of $2\omega_\phi$ as one might guess
596: from, for example, the waveforms computed by Peters
597: and Mathews using Newtonian orbits and the quadrupole formula \cite{peters mathews}.
598: The remaining power is distributed predominantly among three families of modes fixing the integer multipliers
599: for the radial and azimuthal frequencies to be either zero or one. The frequencies for those mode families are
600: \begin{equation}
601: \label{simple families}
602: \omega = \left\{
603: \begin{array}{ll}
604: m\omega_\phi & \text{ dominant power at } m=2 \\
605: m\omega_\phi + \omega_r & \text{ dominant power at } m=2 \\
606: m\omega_\phi + \omega_\theta & \text{ dominant power at } m=1
607: \end{array}
608: \right. ~.
609: \end{equation}
610:
611: From Fig.~\ref{simple spectra}, one can see that the power in any given mode falls off exponentially
612: with $m$, at a rate that is determined by both the orbit geometry, and the values of $k$ and $n$ that define
613: the family. These mode families turn out to dominate the spectra for all orbits with sufficiently small eccentricity
614: and inclination, and the exponential falloff for power in modes within a fixed family turns out to be a general trend
615: for these simple spectra.
616:
617: The distribution of power among modes or mode families is determined by the orbital geometry.
618: The two panels of Fig.~\ref{simple spectra} show that increasing
619: orbital eccentricity draws more power into the family involving the radial frequency, defined by
620: modes (\ref{frequencies}) with $(m, k, n) = (m, 0, 1)$.
621: The spectrum from the system with higher orbital eccentricity also has the greater number of excited
622: modes which are not members of the three families that dominate simple orbits.
623: The following two sections will demonstrate that, in general, the complexity of the spectrum from any EMRI snapshot,
624: or the number of modes excited by any fraction
625: of the total power, is more sensitively dependent on eccentricity than on inclination or
626: semilatus rectum.
627:
628: This general rule that eccentricity governs spectral complexity is in accordance with the preliminary investigation
629: of spectral dependence on orbit geometry given in Ref.~\cite{drasco hughes 2006}. There, significant
630: waveform ``voices'' were defined by sets of frequencies $\omega_{mkn}$ defined as follows
631: \begin{subequations} \label{voices}
632: \begin{align}
633: &\text{azimuthal voice: }& k = 0 \text{ and } n = 0~,\\
634: &\text{polar voice: }& k \ne 0 \text{ and } n = 0~,\\
635: &\text{radial voice: }& k = 0 \text{ and } n \ne 0~,\\
636: &\text{mixed voice: }& k \ne 0 \text{ and } n \ne 0~.
637: \end{align}
638: \end{subequations}
639: For the spectra computed in Ref.~\cite{drasco hughes 2006}, the distribution of power among these voices
640: was more strongly dependent on eccentricity than on the other orbital parameters. From the spectra
641: in Fig.~\ref{simple spectra}, one might guess that these sets of frequencies are the best spectral classification
642: scheme. The most significant of the mode families dominating simple spectra is exactly the azimuthal voice,
643: and the other two dominant mode families are given by one member of either the radial or polar voices.
644: For less simple orbits though, the voices defined above will prove a poor means of classifying spectra.
645: For most generic orbit geometries, the bulk of the power is carried by the mixed voice, and there will prove to be
646: a simple way of grouping the different members of that very large collection of modes.
647:
648: The dominant lines of a third sample spectrum (also computed to a fractional accuracy
649: of $\varepsilon_\text{flux} = 10^{-6}$, in the total radiated power $\left<dE/dt\right>$) is shown in
650: Fig~\ref{hard spectrum}.
651: \begin{figure}
652: \includegraphics[width = .459\textwidth]{Spectrum_3.eps}
653: \caption{\label{hard spectrum} The dominant portion of the spectrum from a highly eccentric, and
654: highly inclined, orbital configuration. The parameters describing the system (black hole spin $a$,
655: eccentricity $e$, semilatus rectum $p$, and inclination $\iota$) are again shown at the top of the plot.
656: For this spectrum, 4\% of the total power is carried by the dominant mode, 99\% is carried by the top 726 modes, and about
657: 50\% is carried by the lines shown. Here the highlighted
658: mode families have frequencies $\omega_{11n} = \omega_\phi + \omega_\theta + n\omega_r$
659: (solid magenta line), $\omega_{02n} = 2\omega_\theta + n\omega_r$ (dashed cyan line),
660: $\omega_{20n} = 2\omega_\phi + n\omega_r$ (dash-dot mustard-colored line), for various values of $n$.
661: All but two of the shown lines are contained in these families (the two excluded lines are drawn slightly
662: thinner than the others).}
663: \end{figure}
664: The orbit for this spectrum is both highly eccentric and highly inclined. The frequencies of
665: the dominant modes are also not given by Eq.~(\ref{simple families}), but are instead
666: \begin{equation}
667: \label{hard families}
668: \omega = \left\{
669: \begin{array}{ll}
670: \omega_\phi + \omega_\theta + n\omega_r & \text{ dominant power at } n=12 \\
671: 2\omega_\theta + n\omega_r & \text{ dominant power at } n=11 \\
672: 2\omega_\phi + n\omega_r & \text{ dominant power at } n=14
673: \end{array}
674: \right. ~.
675: \end{equation}
676: These mode families are not easily classified by the voices (\ref{voices}) of Ref.~\cite{drasco hughes 2006}.
677: The values $n_{\max}$ of $n$ for the dominant members of these mode families can be crudely approximated
678: (to within about 20\% to 40\% for the three families highlighted in Fig.~\ref{hard spectrum}) using the
679: conjecture (\ref{PM peaks}). Equation (\ref{PM peaks}) is a good approximation to the peaks in the spectra derived by Peters and Mathews \cite{peters mathews}
680: \begin{align}\label{PM power}
681: \left\langle \frac{dE}{dt} \right\rangle^{\text{PM}}_{\hat n}\!\!
682: &\propto \frac{{\hat n}^4}{32} \biggl\{ \bigl[ J_{{\hat n}-2}({\hat n}e) - 2eJ_{{\hat n}-1}({\hat n}e) &
683: \nonumber \\
684: & + \frac{2}{{\hat n}}J_{\hat n}({\hat n}e)+2eJ_{{\hat n}+1}({\hat n}e)-J_{{\hat n}+2}({\hat n}e) \bigr]^2&
685: \nonumber \\
686: &+ (1-e^2)\left[ J_{{\hat n}-2}({\hat n}e) - 2J_{\hat n}({\hat n}e)+J_{{\hat n}+2}({\hat n}e) \right]^2&
687: \nonumber \\
688: & + \frac{4}{3{\hat n}^2}\left[ J_{\hat n}({\hat n}e) \right]^2 \biggr\} ,&
689: \end{align}
690: where here $\hat n$ is the multiplier of the single
691: frequency of a Newtonian orbit with eccentricity $e$, and $J_{\hat n}(x)$ are Bessel functions of the first kind.
692: Though a very crude estimator for the values of $n$ describing the dominant members of various mode families, this
693: formula is better than any other estimator that has been tried in the present work.
694:
695: By comparison to the simple spectra, the number of modes needed to capture any fraction of the total power
696: from the complicated spectrum is much larger. This is demonstrated more clearly by Fig.~\ref{falloff}.
697: \begin{figure}
698: \includegraphics[width = .459\textwidth]{falloff.eps}
699: \caption{\label{falloff} The normalized mode amplitudes for the orbital configurations whose corresponding spectra are shown
700: in Figs.~\ref{simple spectra} and \ref{hard spectrum}, sorted in order of decreasing power. The parameters describing the
701: orbital geometry (eccentricity $e$, semilatus rectum $p$, and inclination $\iota$) are shown next to each curve. The crosshairs
702: shown on the solid curve indicate the location of the mode with frequency $\omega = 2\omega_\phi$. That same mode is
703: the dominant mode for both of the other systems, as can be read off of Fig.~\ref{simple spectra}. The inset plot shows the same
704: curves, but on a log-linear scale over a shorter range.}
705: \end{figure}
706: There one can see that, when the modes of a given spectra are sorted by mode number $\Lambda$ in order of decreasing power, mode power falls off roughly as a power law $\Lambda^\alpha$, where $\alpha$ is orbit dependent, for the simple orbits. For the two simple orbits, the
707: rate of this power-law falloff is faster for the less eccentric orbit. If there is a power-law falloff for the more complicated orbit, it is not
708: evident in the first thousand modes. However, given that the essentially blind algorithm from Ref.~\cite{drasco hughes 2006} computed $\approx$ 30,000 modes before finding the dominant thousand or so shown for the complicated orbit in Fig.~\ref{falloff}, most of the total power is still captured by a surprisingly small number of modes.
709:
710: Note also that the mode which dominates the simple orbits, with frequency $2\omega_\phi$, hardly contributes to the spectrum
711: of the complicated orbits. In Fig.~\ref{falloff}, the location of that mode on the curve for the complicated orbit is indicated with a crosshairs at about $\Lambda = 600$, and its power relative to the total power $\approx 6 \times10^{-5}$, is effectively
712: insignificant. This is an extreme example of an effect that has been noticed in simulations of black hole binary inspirals with mass ratios near unity.
713: In the language of post-Newtonian descriptions of such systems (for a recent overview see Ref.~\cite{van den broeck and sengupta}) all modes other than the one with frequency $2\omega_\phi$
714: are ``higher harmonics.'' The excitation of higher harmonics in those systems has been found to be significant
715: in both analytic parameter estimation studies for LISA \cite{sintes vecchio, hellings moore, arun et al, arun et al 2, trias sintes} and fully relativistic numerical simulations \cite{berti et al 2007, vaishnav et al}.
716: LISA parameter estimation studies have to date included either spin precession effects \cite{cutler,vecchio,berti buonanno will,lang hughes} or higher harmonics \cite{hellings moore, arun et al, trias sintes}, but not yet both.
717:
718: \section{Survey of many spectra}\label{Survey of many spectra}
719:
720: In this section, I simulate EMRI snapshots for a large grid of orbital parameters and discuss how the spectral trends
721: identified in the previous section vary over the grid. For each snapshot in the survey, the spin of the
722: large black hole is taken to be $a = 0.8M$, and the spectra are computed with the code described in Ref.~\cite{drasco hughes 2006}
723: using a requested fractional accuracy of $\varepsilon_\text{flux} = 1\%$, in the total radiated power $\left<dE/dt\right>$. The grid of
724: orbital parameters is uniformly spaced in eccentricity $e$,
725: inclination\footnote{It might be more natural to use a uniform distribution in $\cos \iota$.
726: However, this would only be a small effect on the actual values of inclination used. For example, the prograde orbits had ten inclinations uniformly distributed in $\iota$. If ten equally spaced values of
727: $\cos \iota$ were used instead, the average difference in $\iota$ for any of the orbits would have been about $3^\circ$, with the maximum difference being about $6^\circ$. These changes would not significantly affect any of the conclusions in this work.}
728: $\iota$, and in the ratio of the semilatus rectum $p$ to its value for
729: the innermost stable circular orbit (ISCO) $p_{\text{ISCO}}$. The specific values for these parameters are given in Table \ref{OrbitTable}.
730: \begin{table}
731: \begin{tabular}{l l l | l l l}
732: $e$ & $p$ & $\iota$ & $e$ & $p$ & $\iota$ \\ \hline
733: 0.1 & 3.4880 & 20$^\circ$ & 0.1 & 10.118 & 110$^\circ$ \\
734: 0.1444 & 3.7632 & 25.556$^\circ$ & 0.1444 & 10.917 & 115.56$^\circ$ \\
735: 0.1889 & 4.0388 & 31.111$^\circ$ & 0.1889 & 11.716 & 121.11$^\circ$ \\
736: 0.2333 & 4.3140 & 36.667$^\circ$ & 0.2333 & 12.514 & 126.67$^\circ$ \\
737: 0.2778 & 4.5893 & 42.222$^\circ$ & 0.2778 & 13.313 & 132.22$^\circ$ \\
738: 0.3222 & 4.8648 & 47.778$^\circ$ & 0.3222 & 14.112 & 137.78$^\circ$ \\
739: 0.3667 & 5.1401 & 53.333$^\circ$ & 0.3667 & 14.911 & 143.33$^\circ$ \\
740: 0.4111 & 5.4157 & 58.889$^\circ$ & 0.4111 & 15.710 & 148.89$^\circ$ \\
741: 0.4556 & 5.6909 & 64.444$^\circ$ & 0.4556 & 16.509 & 154.44$^\circ$ \\
742: 0.5 & 5.9662 & 70$^\circ$ & 0.5 & 17.307 & 160$^\circ$ \\
743: 0.54 & 6.2417 & & 0.54 & 18.106 & \\
744: 0.58 & 6.5170 & & 0.58 & 18.905 & \\
745: 0.62 & 6.7922 & & 0.62 & 19.703 & \\
746: 0.66 & 7.0678 & & 0.66 & 20.503 & \\
747: 0.7 & 7.3431 & & 0.7 & 21.301 & \\
748: 0.74 & 7.6186 & & 0.74 & 22.101 & \\
749: 0.78 & 7.8939 & & 0.78 & 22.899 & \\
750: 0.82 & 8.1691 & & 0.82 & 23.698 & \\
751: 0.86 & 8.4447 & & 0.86 & 24.497 & \\
752: 0.9 & 8.7199 & & 0.9 & 25.295 &
753: \end{tabular}
754: \caption{\label{OrbitTable}
755: The parameters characterizing the 8000 orbit geometries considered (all about a black hole with spin $a = 0.8M$).
756: The prograde ($\iota < 90^\circ$) are characterized by the 4000 possible combinations of the
757: first three columns, and the retrograde ($\iota > 90^\circ$) are characterized by the 4000 possible combinations of the
758: last three columns. The two different ranges for $p$ correspond to one uniform range for $p / p_\text{ISCO}$.
759: See Fig.~\ref{OrbitPlot} for a graphical representation of the stable orbits.
760: Note that, for prograde ($\iota < 90^\circ$) orbits, $p_\text{ISCO} \approx 2.9066$.
761: Retrograde ($\iota > 90^\circ$) orbits have a less relativistic $p_\text{ISCO} \approx 8.4318$.}
762: \end{table}
763: Of the 8000 orbit geometries shown in Table \ref{OrbitTable},
764: 728 are unstable. For the unstable orbits, the derivative of the radial potential
765: is negative at the prescribed minimum radius
766: \begin{equation}
767: \left. \frac{d}{dr} \left(\frac{dr}{d\lambda}\right)^2 \right|_{r_{\min}} < 0 ~,
768: \end{equation}
769: where $\lambda$ is Mino's time parameter, related to proper time $\tau$ by
770: $d\tau = (r^2 + a^2 \cos^2\theta)d\lambda$, and where the radial potential is
771: \begin{equation}
772: \left(\frac{dr}{d\lambda}\right)^2 = \left( E\varpi^2 - a L \right)^2
773: - \Delta\left[\mu^2 r^2 + (L - a E)^2 + Q\right]~.
774: \end{equation}
775: Here $\varpi^2 = r^2 + a^2$ and $\Delta = r^2 - 2Mr + a^2$.
776: Snapshot spectra were computed for each of the remaining 7272 stable orbital configurations,
777: represented graphically in Fig.~\ref{OrbitPlot}. The total computational cost of simulating these spectra
778: was about 2.7 CPU-years on a machine based on a 3.2 GHz Intel Pentium 4 Xeon processor.
779: \begin{figure}
780: \includegraphics[width = .459\textwidth]{Orbits_stable_rmin.eps}
781: \caption{A graphical representation of the stable orbits described by the parameters shown in Table
782: \ref{OrbitTable}.
783: The inclination and minimum radius ($r_{\min}$) for the 7,272 stable orbits are shown.}
784: \label{OrbitPlot}
785: \end{figure}
786:
787: All of the spectra from this grid are dominated by mode families characterized by
788: frequencies
789: \begin{equation}
790: \label{top two families}
791: \omega = \left\{
792: \begin{array}{l}
793: \pm2\omega_\phi + n\omega_r \\
794: \pm\omega_\phi + \omega_\theta + n\omega_r
795: \end{array}
796: \right. ~,
797: \end{equation}
798: where $\pm$ was 1 for prograde orbits and $-1$ for retrograde orbits.
799: To within an error on the order of 10\%, the most dominant mode for either family
800: had $n = n_{\max}$, where $n_{\max}$ is given by the Peters-Mathews approximation (\ref{PM peaks}).
801: The success of that approximation remains similar to that for the more complicated
802: sample spectrum in Fig.~\ref{hard spectrum}, and is plotted for the entire set of orbits
803: in Fig.~\ref{nmax vs e}.
804: \begin{figure}
805: \includegraphics[width = .459\textwidth]{nmax_vs_e.eps}
806: \caption{The circles indicate the value of $n$ in $\omega_{mkn}$ for the most powerful
807: modes in the spectra from the stable orbits in Table \ref{OrbitTable}.
808: The solid curve is the approximation (\ref{PM peaks}) that gives the peaks in the
809: approximate spectra (\ref{PM power}) derived for Newtonian orbits by Peters and
810: Mathews \cite{peters mathews}.}
811: \label{nmax vs e}
812: \end{figure}
813: Of the two dominant mode families (\ref{top two families}) the
814: first one, with $(m, k, n) = (2, 0, n)$, is the most common. The exceptions, dominated by the $(1, 1, n)$-family,
815: are the orbits with $\iota$ nearest to 90$^\circ$. This trend is demonstrated
816: in Table \ref{elevens}.
817: \begin{table}
818: \begin{tabular}{c c}
819: $\iota$ & Fraction of orbits \\ \hline
820: 20$^\circ$ & 0\\
821: 25.556$^\circ$ & 0\\
822: 31.111$^\circ$ & 0\\
823: 36.667$^\circ$ & 0\\
824: 42.222$^\circ$ & 0\\
825: 47.778$^\circ$ & 0\\
826: 53.333$^\circ$ & 0\\
827: 58.889$^\circ$ & 26\% \\
828: 64.444$^\circ$ & 94\% \\
829: 70$^\circ$ & 100\%\\
830: 110$^\circ$ & 100\%\\
831: 115.56$^\circ$ & 100\%\\
832: 121.11$^\circ$ & 100\%\\
833: 126.67$^\circ$ & 100\%\\
834: 132.22$^\circ$ & 1.8\%\\
835: 137.78$^\circ$ & 0.3\%\\
836: 143.33$^\circ$ & 0.3\%\\
837: 148.89$^\circ$ & 0.3\%\\
838: 154.44$^\circ$ & 0.3\%\\
839: 160$^\circ$ & 0.3\%
840: \end{tabular}
841: \caption{\label{elevens}
842: The fraction of orbits, from the grid of orbits described in Table \ref{OrbitTable},
843: dominated by a mode with frequency $\omega_{mkn}$,
844: where $(m, k, n) = (\pm1, 1, n)$, as a function of inclination $\iota$.
845: The remaining orbits are dominated by modes with $(m, k, n) = (\pm2, 0, n)$.
846: Here $\pm$ is $1$ for prograde orbits ($\iota < 90^\circ$) and $-1$ for retrograde orbits ($\iota > 90^\circ$). Note that since
847: the grid of orbits is uniformly spaced in $p/p_\text{ISCO}$, rather than in $p$, the
848: prograde orbits have smaller values for $p$ than do the retrograde orbits.}
849: \end{table}
850: The prograde orbits tend to be less easily dominated by the $(1, 1, n)$ family.
851: Since the orbit grid is evenly spaced in $p/p_{\text{ISCO}}$, and since $p_{\text{ISCO}}$ is much smaller
852: for prograde orbits than for retrograde orbits, this suggests that the closer the orbit comes
853: to the horizon, the harder it becomes for the system to channel radiation away from the $(2, 0, n)$-family and
854: into the $(1, 1, n)$ family.
855:
856: As with the sample orbits, the number of modes needed to capture the bulk of the radiated power remains
857: resonably small. This is shown in Fig.~\ref{N99 histogram} by a histogram of the number of modes $N_{99\%}$ carrying 99\% of the power.
858: \begin{figure}
859: \includegraphics[width = .459\textwidth]{N99_histogram.eps}
860: \caption{A histogram showing the fraction of orbits, out of the 7,272 stable orbits in
861: Table \ref{OrbitTable}, for which $N_{99\%}$ modes are required to capture 99\% of the
862: radiated power. The inset displays the same data as a cumulative distribution (its vertical
863: axis showing the fraction of orbits with $N_{99\%}$ less than the value on the horizontal axis).}
864: \label{N99 histogram}
865: \end{figure}
866: An example of the dependence of $N_{99\%}$ on orbit geometry is plotted in Fig.~\ref{N99 contour}.
867: \begin{figure}
868: \includegraphics[width = .459\textwidth]{N99_contour_color.eps}
869: \caption{The number of modes required to capture 99\% of the
870: radiated power, $N_{99\%}$, as a function of eccentricity $e$ and inclination $\iota$
871: for a fixed value of $p/p_{\text{ISCO}}$. Since $p_{\text{ISCO}}$ is larger for
872: retrograde orbits, the top block of this plot corresponds to orbits with larger values of
873: $p$.}
874: \label{N99 contour}
875: \end{figure}
876: That plot shows the values of $N_{99\%}$ for all the orbits with the smallest value of $p/p_{\text{ISCO}}$,
877: such that over the grid's range for the other orbital parameters, no orbits were unstable.
878:
879: It is important to emphasize that the values of $N_{99\%}$ given in this and the following section
880: are accurate only to $\sim 10\%$. This is because $N_{99\%}$ is found only after the algorithm from Ref.~\cite{drasco hughes 2006}
881: computes a much larger number of modes $N$ in an effort to determine the total power to its requested fractional accuracy
882: of $\varepsilon_\text{flux} = 1\%$. Once that algorithm terminates, the modes that it computed are sorted in order of decreasing
883: power such that
884: \begin{equation}
885: \left< \frac{dE}{dt} \right> = \sum_{\Lambda = 1}^N \left< \frac{dE}{dt} \right>_\Lambda~,
886: \end{equation}
887: where $\left<dE/dt\right>_\Lambda$ decreases with $\Lambda$. The value of $N_{99\%}$ is the smallest possible
888: value satisfying
889: \begin{equation}
890: \sum_{\Lambda = N_{99\%}}^N \left< \frac{dE}{dt} \right>_\Lambda < (0.01) \left< \frac{dE}{dt} \right>~.
891: \end{equation}
892: Spot checking its dependence on $\varepsilon_\text{flux}$ for a few sample spectra gave an expected accuracy $\sim 10\%$.
893:
894: Figure \ref{N99 contour} also shows that, as with the sample orbits, $N_{99\%}$ is more strongly dependent on eccentricity
895: than on inclination and semilatus rectum. This is especially significant for the prospect of observing intermediate mass ratio inspirals with ground-based detectors like LIGO, since the systems that have been estimated to be the most likely candidates for
896: being observed are those with especially small eccentricity, typically $e < 10^{-4}$ and at most $e\approx 0.1$ \cite{mandel et al 2007}. For LIGO, it is likely that waveforms needed for detection need only
897: contain a few to $\sim 10$ modes. This both simplifies LIGO's task of detection, since the waveform snapshots will not be very complicated, and hardens its task of spacetime mapping, since correspondingly less information will be observable.
898:
899:
900:
901: \section{Spectra from a kludged inspiral}\label{Spectra from a kludged inspiral}
902:
903: In this section, I describe how general spectral characteristics should be expected to evolve during an EMRI.
904: The snapshots studied here will be sampled at approximately 12-hour intervals over the final three years of a
905: single kludged EMRI thought to be typical of the kind that could be observed by LISA.
906: The kludged trajectory through orbital parameter space $e(t)$, $\iota(t)$, and $p(t)$, was
907: provided by Jonathan Gair, and was numerically computed according to the prescription introduced in
908: Ref.~\cite{gair glampedakis 2005}.
909: Their method for approximating the trajectory is based on an eclectic combination of approximations
910: including post-Newtonian equations
911: for the radiative fluxes of energy and angular momentum, numerical fits to Teukolsky-based calculations of fluxes for circular
912: and equatorial orbits, as well as some uncontrolled approximations for the evolution of Carter's constant. Since
913: only power spectra will be discussed here, the results are independent of any evolution for the positional orbital elements, or
914: $\chi_{mkn}$ in Eq.~(\ref{positional phase}).
915: While one would expect minimal accuracy from such an array of approximations, these kludged trajectories
916: have been shown to exhibit a stunning degree of agreement with more accurate calculations.
917: The integrated overlap between approximate waveforms based on the kludged orbit trajectories
918: and waveforms constructed from black hole perturbation theory alone is often about 95\% \cite{babak et al 2007}.
919: So a kludged orbit trajectory is likely more than adequate for the present purpose, since the snapshots
920: themselves will still be accurate up to leading order in the mass ratio, and since for the majority of the spectra examined in
921: the previous section, the general character of a spectrum does not change dramatically for small perturbations to the orbital
922: parameters.
923:
924: The kludged orbit trajectory for the inspiral examined here is shown in Fig.~\ref{kludge}\footnote{
925: There is no special reason for using a black hole spin of $a = 0.9M$ in the kludged trajectory
926: as opposed to $a = 0.8M$, as is used for the grid of orbits in the previous section. The
927: difference is due to the trajectory and the present work not being produced in parallel.}
928: \begin{figure}
929: \includegraphics[width = .459\textwidth]{frequencies.eps}
930: \includegraphics[width = .459\textwidth]{KITS_e.eps}
931: \caption{\label{kludge} The kludged evolution for a selection of the orbits principal
932: elements during an inspiral of the sort that could be observed with LISA. For this system, the
933: large black hole has mass $M = 10^6 M_\odot$, and the magnitude
934: of its spin angular momentum is $a = 0.9M$.
935: The smaller black hole is nonspinning and has mass $\mu = 10 M_\odot$.
936: The initial orbit has eccentricity $e = 0.79$, semilatus rectum $p = 8.8$, and inclination $\iota = 43^\circ$.
937: The final orbit has eccentricity $e = 0.10$, semilatus rectum $p = 3.2$, and inclination $\iota = 45^\circ$.
938: The orbit trajectory $e(t)$, $p(t)$, and $\iota(t)$ for this inspiral was computed by Jonathan Gair using the
939: algorithm described in Ref.~\cite{gair glampedakis 2005}. The approximations used to compute
940: the trajectory should become less accurate with time as the system becomes increasingly relativistic.
941: The top panel shows the evolution of the orbital frequencies $f_{r,\theta,\phi} = \omega_{r,\theta,\phi}/(2\pi)$
942: with an inset of the same but over a different range. The bottom panel shows the evolution of orbital eccentricity $e$,
943: with an inset of the same but over a different range.}
944: \end{figure}
945: The principal orbital elements evolve slowly and smoothly throughout the majority of the inspiral.
946: In the final few days of the inspiral, where the trajectory should be least accurate and where the
947: adiabatic approximation itself should begin to fail, the eccentricity begins to rise with time, and the orbital
948: frequencies rapidly diverge from each other. The trajectory ends when it has evolved onto an unstable orbit,
949: at which point the binary would merge to form a single black hole.
950:
951: Snapshot spectra were computed at roughly 12-hour
952: intervals along the orbit trajectory in Fig.~\ref{kludge}.
953: As was done for the grid of spectra in the previous section, these spectra were computed with the code described in
954: Ref.~\cite{drasco hughes 2006} using a requested
955: fractional accuracy of $\varepsilon_\text{flux} = 1\%$, in the total power $\left<dE/dt\right>$.
956: There were 2,143 snapshots in all, and the total computational
957: cost of simulating them was about 1.5 CPU-years on a machine based on a 3.2 GHz Intel Pentium 4 Xeon processor.
958: The initial and final snapshot spectra from this sequence are shown in Fig.~\ref{first and last spectra}.
959: \begin{figure}
960: \includegraphics[width = .459\textwidth]{FirstLast.eps}
961: \caption{\label{first and last spectra} The initial (blue, or dark grey) and final (red, or light grey) power spectra for the inspiral
962: shown in Fig.~\ref{kludge}. An animated movie of the complete spectral evolution is available upon request.}
963: \end{figure}
964: The inspiral's initial spectrum is similar in character to the one shown in Fig.~\ref{hard spectrum}, due to the large initial orbital
965: eccentricity. The final spectrum is more similar to the ones in Fig.~\ref{simple spectra}.
966: The dominant mode families for both the initial and final spectra have $(m, k, n) = (2, 0, n)$
967: and $(1, 1, n)$. Those two families make up the two largest arcs or lobes of lines in the initial spectrum.
968: In the final spectrum, the lobes are much more narrow, and are not as easily identified by eye. For the final spectrum,
969: a more efficient definition of mode families might instead use $m$ as the free index, rather than $n$. For example,
970: the strongest lines along the upper right edge of the final spectrum are $(m, 0, m)$, peaked at $m = 2$.
971:
972:
973: Figure \ref{KITS N99} shows the evolution in spectral complexity during the inspiral.
974: \begin{figure}
975: \includegraphics[width = .459\textwidth]{KITS_N99.eps}
976: \caption{\label{KITS N99} The evolution of the number of modes needed to capture 99\% of the
977: radiated power $N_{99\%}$ for the inspiral shown in Fig.~\ref{kludge}. The inset shows a close-up view
978: over about the last week of the inspiral.}
979: \end{figure}
980: As the binary becomes more circular, the number of lines carrying 99\% of the spectral power
981: $N_{99\%}$ decreases with time until the last few days,
982: at which point both the kludge and adiabatic approximations should fail. Over those last few days
983: the orbital eccentricity rises and $N_{99\%}$ jumps from about 125 to 250.
984: This effect is typical of both kludged trajectories and the more accurate Teukolsky-based
985: trajectories. While it may be a physical effect, it always occurs suspiciously in regimes where the adiabatic
986: approximation and kludge should be least accurate. The question of whether or not it is physical
987: might best be addressed by the numerical relativity community. The causes of the smaller
988: abrupt jumps in $N_{99\%}$ (for example, at either end of a six month gap centered at about one year)
989: is unknown, however, they are within the expected accuracy $\sim 10\%$ whereas the final jump from 125 to 250 is not.
990: Repeating the snapshot calculations with a smaller requested overall accuracy in the total power $\varepsilon_{\text{flux}}$
991: would likely eliminate most, if not all, of these small jumps.
992:
993: \section{Verification of black hole orbits}\label{Verification of black hole orbits}
994:
995: In this section I suggest EMRI detection algorithms based on the spectral trends found above and
996: discuss the scientific meaning of successful detections. The tone of the discussion is meant to be exploratory
997: rather than exhaustive. That is, it is meant to outline the sorts of data analysis strategies that the
998: simulations described above suggest would be useful, rather than to look in great detail at any one algorithm.
999:
1000: This section will be divided up into two subsections. The first subsection outlines algorithms that could search for
1001: systems without radiation reaction. These are really just of interest as building blocks for more complicated searches
1002: since only a small subset of EMRIs would be observable without accounting for the influence of radiation on the binary.
1003: The second subsection describes how these building blocks would be used to search for realistic systems affected by
1004: radiation reaction and gives a detector-independent estimate of how well these algorithms might
1005: perform in the best of circumstances.
1006:
1007: \subsection{Without radiation reaction}
1008:
1009: For times that are sufficiently short for the orbit of the captured mass to be unaffected by radiation\footnote{
1010: For a subset of observable systems ~\cite{drasco 2006} these ``short'' times are actually longer than the longest amounts of time for which
1011: fully coherent search algorithms are computationally affordable (a few weeks \cite{gair et al 2004}).
1012: }, the general expression for EMRI snapshots (\ref{waveform model}) is accurate up to corrections that are at most
1013: of order $\mu^2/M^2$ and are perhaps even as small as $\sim \mu^3/M^3$ with the principal frequencies adjusted according
1014: to the second order metric perturbation \cite{mino 2007}. In this subsection, I will outline a detection strategy for a
1015: signal of this form. Truncating the general expression for an EMRI-snapshot waveform (\ref{waveform model}) to a finite number
1016: of $N$ modes, rewriting it as a single sum, and explicitly showing the positional phase elements, gives the following
1017: \begin{eqnarray}\label{truncated model}
1018: h_{+} - i h_{\times} &=&
1019: \sum_{\Lambda=1}^N h_{\Lambda} \exp\left\{-i [m_\Lambda \omega_\phi (t-t_\phi) \right. \nonumber \\
1020: &+&\left. k_\Lambda \omega_\theta (t-t_\theta)+ n_\Lambda \omega_r (t-t_r)]\right\}~.
1021: \end{eqnarray}
1022: I wish to consider the prospects of using this truncated expression (\ref{truncated model}) explicitly as a
1023: phenomenological template.
1024:
1025: By phenomenological templates, I mean templates for which some of the
1026: free and measurable parameters will have no immediately obvious physical meaning. Traditional, or nonphenomenological
1027: templates are constructed as follows. You start by declaring the 17 parameters which completely determine
1028: the waveform (e.g.~the two masses, and 3-vectors for the position and velocity of each object as well as for the
1029: spin of the larger object). You then solve Einstein's equation or some perturbed form of it to
1030: completely determine $h_{+,\times}$. This last step is equivalent to determining all of the frequencies $\omega_{r, \theta, \phi}$,
1031: phase shifts $t_{r, \theta, \phi}$, and Fourier coefficients $h_{\Lambda}$ in the model (\ref{waveform model}).
1032: The phenomenological templates are computed in a much simpler way. You first chose values for the frequencies, phase shifts, and
1033: Fourier coefficients, and then you simply evaluate the sum over modes (\ref{truncated model}).
1034: For these waveforms, the frequencies, phase shifts, and Fourier coefficients are themselves the free quantities
1035: to be measured. For a phenomenological template with $N$ modes, there are then $5N+6$ free parameters. Of those there are
1036: $2N+6$ real numbers (the three frequencies, the three phase shifts, and the $N$ complex mode amplitudes), as well as $3N$
1037: integers (the frequency multipliers $m_\Lambda$, $k_\Lambda$, and $n_\Lambda$).
1038:
1039: The most obvious objection to the use of these phenomenological templates in gravitational wave searches
1040: is the large dimensionality of the parameter space.
1041: For templates that allow for any more than $N=2$ modes, you will have more free parameters than the
1042: traditional templates\footnote{The specific value $N=2$ here is not exact. The three integer parameters are
1043: much simpler degrees of freedom from the standpoint of a search (discrete and reasonably confined).
1044: In the same vein though, not all of the 17 traditional parameters are intrinsic. So while the value of $N$
1045: for which both the traditional and phenomenological template spaces are dimensionally equivalent is surely small,
1046: the exact value is not obvious.}, and of those only the three
1047: frequencies $\omega_{\phi,\theta,r}$ will carry immediate physical meaning.
1048: In the traditional scheme, however, the step of going from the complete set of 17 parameters to the waveform is
1049: extremely costly. Since both template families are equally good matches to the true waveforms, and since either way you will be dealing
1050: with a significant number of template parameters, it may be worth adding dimensionality to the template parameter space in exchange
1051: for not having to solve Einstein's equation.
1052:
1053: Though the large dimensionality may seem daunting, especially for those familiar with efforts focused on
1054: sources with just a few degrees of freedom, it is not unrealistic for elaborate algorithms to identify signals characterized by
1055: a large number of parameters. For algorithms that match against templates over a sequence of increasingly dense grids on the model
1056: parameter space, computational cost grows as a power law where the power is proportional to the number of free parameters. Algorithms
1057: designed for more complex models have costs that instead grow only linearly with the number of parameters. Algorithms of this nature have
1058: been used in the mock LISA data challenges \cite{mldc} to recover $\sim 10^4$ white dwarf binaries.
1059:
1060: I now discuss how well phenomenological detection algorithms can be expected to perform in the best case scenario
1061: where the algorithm has no difficulty in selecting the correct parameter values. This can be done without reference
1062: to specific instruments, and in a sky averaged sense, by studying the distribution of the power among various
1063: modes relative to the total power radiated by the binary. More sophisticated estimates that account for detector characteristics
1064: and variation of signal parameters are certainly possible, but they are beyond the scope of this paper.
1065:
1066: For the task of searching only for EMRI snapshots with phenomenological templates, the example spectra from Sec.~\ref{Sample spectra}
1067: suggest estimates of the minimum scale of the phase space dimensionality. For example, systems with eccentricities of
1068: about 1\% will produce spectra similar to Fig.~\ref{simple spectra}. So a ground-based phenomenological search for these systems would
1069: require $N\lesssim 10$ modes, resulting in a model with 56 parameters (26 real numbers and 30 relatively small integers).
1070: For snapshots thought to be typical of EMRIs that could be observed with space-based detectors, one would need $N\sim10^2$ to $10^3$.
1071: This would mean a model with about 500 to 5000 free parameters, still far fewer than what has been demonstrated already for white
1072: dwarf binaries \cite{mldc}.
1073:
1074: If a search for phenomenological EMRI snapshots were successful, only the general waveform model would be verified.
1075: One would only be able to claim detection of some signal with a discrete triperiodic
1076: spectrum with some measured fundamental frequencies $\omega_{\phi, \theta, r}$, since the physical meaning of the other unknown
1077: parameters is convoluted. In this event, one could then turn to the underlying physical model. Finding a set of its 17
1078: free parameters that best reproduce the parameters measured in the phenomenological search would then confirm
1079: the physical model to some more explicit level of uncertainty. Failing to find parameters for the underlying physical
1080: model might mean that the snapshot was produced by a test mass moving along a geodesic of some non-Kerr spacetime, since
1081: many (but not all) candidates for such orbits are also triperiodic \cite{gair li mandel 2007, flanagan hinderer 2007}.
1082:
1083: Some intermediate level of model verification is also possible and could reduce the dimensionality of the parameter space for the
1084: phenomenological templates. For example, sixteen mode families are needed to capture
1085: 99\% of the power radiated by the initial snapshot from the inspiral examined in the previous section.
1086: The distribution of power among these families is shown in Table \ref{first families}.
1087: \begin{table}
1088: \begin{tabular}{c c c c}
1089: $m$ & $k$ & $n_{\max}$ & $ \left<\dot E\right>_{mk} / \left<\dot E\right>$ \\ \hline
1090: 2 & 0 & 16 & 5.1$\times 10^{-1}$ \\
1091: 1 & 1 & 15 & 2.2$\times 10^{-1}$ \\
1092: 3 & 0 & 27 & 9.2$\times 10^{-2}$ \\
1093: 2 & 1 & 25 & 6.2$\times 10^{-2}$ \\
1094: 0 & 2 & 13 & 3.5$\times 10^{-2}$ \\
1095: 4 & 0 & 38 & 2.0$\times 10^{-2}$ \\
1096: 1 & 2 & 24 & 1.7$\times 10^{-2}$ \\
1097: 3 & 1 & 36 & 1.8$\times 10^{-2}$ \\
1098: 2 & 2 & 35 & 6.4$\times 10^{-3}$ \\
1099: 2 & -1 & 12 & 3.7$\times 10^{-3}$ \\
1100: 4 & 1 & 47 & 2.9$\times 10^{-3}$ \\
1101: 5 & 0 & 49 & 2.5$\times 10^{-3}$ \\
1102: -1 & 3 & 12 & 1.9$\times 10^{-3}$ \\
1103: 0 & 3 & 22 & 1.3$\times 10^{-3}$ \\
1104: 3 & 2 & 45 & 1.5$\times 10^{-3}$ \\
1105: 3 & -1 & 21 & 9.9$\times 10^{-4}$
1106: \end{tabular}
1107: \caption{\label{first families}
1108: The 16 mode families needed to capture 99\% of the power radiated
1109: during the first 12 hours of the inspiral discussed in Sec.~\ref{Spectra from a kludged inspiral}.
1110: Each family is defined by fixing $m$ and $k$. The most powerful member of each family has
1111: frequency $m \omega_\phi + k \omega_\theta + n_{\max} \omega_r$. The ratio of the power
1112: radiated by each family to the total power from this first 12 hours is given
1113: by $\left<\dot E\right>_{mk} / \left<\dot E\right>$. For this system, the spin of the black hole is $a = 0.9M$,
1114: and the orbit of the test mass has eccentricity $e = 0.79$, semilatus rectum $p = 8.8$, and inclination $\iota = 43^\circ$.}
1115: \end{table}
1116: As is true for all the snapshots simulated in this paper, this one is dominated by the mode families with $(m, k, n) = (2, 0, n)$ and
1117: $(1, 1, n)$. And as is typical, those two mode families carry most of the power, 73\% here. In an effort to simplify the
1118: phenomenological waveform model, one might restrict it to include only those mode families.
1119: This specific model would eliminate ($m_{\Lambda}$, $k_{\Lambda}$) from the template parameter space by fixing them
1120: to either $(2, 0)$ or $(1, 1)$, and would create two new parameters specifying the number of modes in each of the two families.
1121: This would reduce the number of free parameters from $5N+6$ to $3N+8$.
1122:
1123: \subsection{With radiation reaction}
1124:
1125: The scenario describe in the previous subsection is only immediately useful for EMRI's with the most extreme
1126: mass ratios. There is no compelling reason to expect those systems to be especially common,
1127: or to even consider them reasonable targets at all. The purpose of studying these simple systems is to
1128: construct from the results an approximate description of more generic EMRIs that respond to their own radiation.
1129: That is, we wish to describe the radiation of a generic adiabatic EMRI as a slowly evolving sequence of EMRI
1130: snapshots. Here I will now outline how the phenomenological templates for EMRI snapshots could be modified to
1131: describe the more general class of adiabatic EMRIs. There are many ways that this could be done.
1132: Although a detailed study of specific models would be valuable, it is beyond the scope of this paper.
1133: Instead, I aim to be as general as possible and will steer away from discussing any
1134: specific implementation.
1135:
1136: For an adiabatic EMRI, the motion of the small object is described by a solution of the geodesic equation for
1137: the Kerr spacetime, but with the orbital elements replaced by quantitates that evolve slowly. The Teukolsky equation
1138: can provide the leading order radiative changes to those quantities, and more sophisticated techniques are envisioned
1139: for describing both conservative and radiative effects. In the spirit of trading calculation difficulty for added dimensionality, every quantity that was constant for the snapshot model (\ref{truncated model}) could in principle be
1140: replaced by simple, one or two parameter models.
1141:
1142: To illustrate this, consider the orbital frequencies $\omega_i$, for $i = r$, $\theta$, $\phi$.
1143: These can be taken to drift linearly with time
1144: \begin{equation}
1145: \omega_i \to \omega_{i} + \dot \omega_{i} t~.
1146: \end{equation}
1147: where I have introduced new constants $\dot \omega_{i}\propto \mu/M$.
1148: This simple model fits the first year of the three frequency
1149: trajectories shown in Fig.~\ref{kludge} with an average fractional accuracy of about 1\%.
1150: Other models can of course do better. For example, Peters and Mathews derived an expression for radius as a function
1151: of time in the case of slow circular inspirals. Combining that result,
1152: Eq.~(5.9) of Ref.~\cite{peters mathews}, with Kepler's law $M\omega = (r/M)^{-2/3}$, gives
1153: \begin{equation}
1154: M\omega = \left[\left(\frac{r_0}{M}\right)^4 - \frac{256}{5}\frac{\mu}{M}\frac{t}{M} \right]^{-3/8}~.
1155: \end{equation}
1156: For the more general case of fast generic motion, one might want to try a model with a similar form
1157: \begin{equation}
1158: M\omega_i = [\alpha_i + \beta_i (t/M)]^{-\gamma_i}~.
1159: \end{equation}
1160: This model fits the first year of the three frequency trajectories shown in Fig.~\ref{kludge} with
1161: $\gamma_i \approx 1.6$ and an average fractional accuracy on the order of $10^{-4}$.
1162: It performs similarly at later times, but not with the same values of the parameters
1163: $\alpha_i$, $\beta_i$, and $\gamma_i$. Unlike the simple linear model however, this one has
1164: three parameters instead of two.
1165:
1166: To complete the construction of phenomenological templates for adiabatic EMRIs,
1167: similar models can be concocted for the other parameters of the snapshot
1168: templates (\ref{truncated model}), the complex mode amplitudes, and the positional phase elements.
1169: If one-parameter models like the linear frequency drift are used, then the new template is given
1170: by Eq.~(\ref{truncated model}) with
1171: \begin{eqnarray}
1172: h_\Lambda &\to& h_\Lambda + \dot h_\Lambda t ~,\\
1173: m_\Lambda &\to& m_\Lambda + \dot m_\Lambda t ~,\\
1174: k_\Lambda &\to& k_\Lambda + \dot k_\Lambda t ~,\\
1175: n_\Lambda &\to& n_\Lambda + \dot n_\Lambda t ~,\\
1176: \omega_i &\to& \omega_{i} + \dot \omega_{i} t~. \\
1177: t_i &\to& t_i + \dot t_i t~,
1178: \end{eqnarray}
1179: again for $i = r$, $\theta$, $\phi$. For this case, the dimensionality of the template parameter space is
1180: doubled to $10 N + 12$. Of those, only the $3N$ frequency multipliers are integers.
1181: These phenomenological templates are similar in nature to the time-frequency search methods which have
1182: been successfully demonstrated in the mock LISA data challenge \cite{gair et al 2007}. They differ
1183: from those in that they can accommodate coherent integration and could also more naturally include
1184: specific schemes for evolving both the principle and positional orbital elements.
1185:
1186: Given that EMRI snapshots tend only to be dominated by a small number of mode families, it is likely that
1187: a slightly simpler waveform model could be used for adiabatic EMRIs.
1188: Figure \ref{captured power} demonstrates how successful two such hypothetical detection algorithms might be
1189: if searching for the kludged inspiral from Sec.~\ref{Spectra from a kludged inspiral}.
1190: \begin{figure}
1191: \includegraphics[width = .459\textwidth]{CapturedPower.eps}
1192: \caption{\label{captured power} A time-dependent measure of success for two hypothetical detection
1193: algorithms searching for the inspiral from Fig.~\ref{kludge}. The algorithm for the dashed blue
1194: curve captures the power radiated at all frequencies of the form $\omega_{mkn}$, where
1195: $(m, k, n) = (1, 1, n)$, $(0, 2, n)$, and $(2, 0, n)$, for any $n$. The solid curve captures power
1196: radiated at all frequencies in the 16 mode families needed to capture 99\% of the initial power,
1197: where here a mode family is defined as all modes with frequencies $\omega_{mkn}$ for some fixed $m$ and $k$.}
1198: \end{figure}
1199: If one were to search for this waveform using a model that included only the dominant three mode families
1200: from the sample complicated snapshot in Fig.~\ref{hard spectrum} one would recover $71\%$ of the inspiral's total power.
1201: Assuming this could be done with one-parameter models for the adiabatic evolution of the waveforms parameters, such a model
1202: would have $6N+18$ free parameters (by eliminating $2N$ frequency multipliers and their $2N$ linear drifts, and by adding
1203: three new integers specifying how many modes are in each family, as well as their three linear drifts).
1204: For this example EMRI, this scheme included $N=121$ modes. So 714 free parameters would have been needed to recover 71\% of the power.
1205: A similar hypothetical algorithm which captured all the power carried by the 16
1206: mode families that make up most of that inspiral's initial spectrum (Table \ref{first families}) would recover 98\% of the total
1207: power with $6N+44$ free parameters (by eliminating $2N$ frequency multipliers and their $2N$ linear drifts, and by adding
1208: 16 new integers specifying how many modes are in each family, as well as their 16 linear drifts).
1209: For this example EMRI, this scheme included $N=411$ modes. So 2,510 free parameters would have been needed to recover 98\% of the power.
1210:
1211: It should be emphasized that any gravitational wave detections following from the use of these phenomenological waveform models would not necessarily yield either the parameters that completely determine an EMRI (position, masses, etc) or the spacetime
1212: map that general relativity predicts is encoded in the radiation. They would however verify the detection of waveforms predicted to
1213: be produced when a test mass perturbs a rotating black hole by moving through an adiabatic sequence of its bound geodesic orbits.
1214: They would also measure the evolution of the three fundamental frequencies throughout that sequence, or equivalently the evolution
1215: for any other set of principle orbital elements. It is possible that restricting phenomenological EMRI waveforms to include only
1216: the mode families that are most commonly dominant in the snapshots simulated here may alone be enough of a constraint to keep EMRIs
1217: into non-Kerr black hole candidates from triggering a detection. However, without more work with the snapshots from such EMRIs
1218: one could not say so with any certainty. One would have to be content only to have verified that radiation from an adiabatic sequence
1219: of black hole orbits could have triggered the detection.
1220:
1221: \section{Conclusion}\label{Conclusion}
1222:
1223: The number of significant modes in generic EMRI-snapshot spectra has been shown generally to be
1224: much more manageable than one might have guessed from earlier truncation
1225: algorithms \cite{drasco hughes 2006}.
1226: This should lead to improved truncation algorithms, which will reduce the cost of future data analysis efforts.
1227: Such improvements should exploit the trends observed here in the relationship between orbit geometry and spectral signature. The ability to predict the multiplier $n \approx n_{\max}$ of the radial frequency for the dominant modes
1228: using a formula based on such simple approximations \cite{peters mathews} is encouraging. It suggests that many
1229: of the trends in these spectra might be understood analytically using more recent
1230: tools \cite{moreno-garrido mediavilla buitrago, poisson, ganz et al}.
1231:
1232: The detection algorithms that are suggested here for verifying minimal aspects of relativity and black hole
1233: physics may ultimately be used in future gravitational wave detections. However, more work is needed to determine
1234: whether or not they are cost-efficient and science-efficient alternatives to traditional search techniques.
1235: Another possibly interesting area for future work is to explore the dependence of the snapshot spectra on the spin of
1236: the larger black hole. It has been implicitly assumed that the large values of spin considered are somehow representative
1237: of observable EMRIs. This is reasonable since the few existing measurements due to modeling x-ray spectra from
1238: galactic nuclei suggest near-maximal spins \cite{miller, reis et al}.
1239: Still, future work that tests the generality of these parameter values by simulating other EMRIs would be worthwhile.
1240:
1241: \begin{acknowledgments}
1242: I am especially grateful to Scott Hughes for providing significant portions of the numerical code used in this work, and to Jonathan Gair for providing the kludged inspiral trajectory.
1243: I would like to thank Chao Li, Geoffrey Lovelace, and Kip Thorne for discussions that triggered this work. I also thank Emanuele Berti, Yasushi Mino, and Michele Vallisneri for encouragement and helpful discussions.
1244: The supercomputers used in this investigation were provided by funding from the JPL Office of the Chief Information Officer.
1245: This research was carried out at the Jet Propulsion Laboratory, California Institute of Technology,
1246: under a contract with the National Aeronautics and Space Administration and funded through the
1247: internal Human Resources Development Fund initiative, and the LISA Mission Science Office.
1248: \end{acknowledgments}
1249:
1250: \appendix
1251: \section{Bi periodic form of time-dependent spatial coordinates}\label{append}
1252:
1253: Here I derive the bi-periodic form (\ref{coordinate orbits}) of the
1254: spatial Boyer-Lindquest coordinates $r$, $\theta$, and $\phi$ for bound
1255: geodesics as a function of the time coordinate $t$.
1256:
1257: The bi-periodic forms of both the radial and polar coordinates follows immediately from
1258: Sec.~IV of Ref.~\cite{drasco hughes 2004} by replacing $f[r(t),\theta(t)]$ with $r(t)$
1259: and $\theta(t)$. The resulting relationships between the coefficients in the Mino-time series
1260: expansion (\ref{mino orbits}) and the coordinate-time expansions (\ref{coordinate orbits}) are given by
1261: \begin{eqnarray}
1262: {\tilde r_{kn}} &=&
1263: \frac{\Upsilon_r \Upsilon_\theta}{(2\pi)^2\Gamma}
1264: \int_0^{2\pi/\Upsilon_r} d\lambda_r
1265: \int_0^{2\pi/\Upsilon_\theta} d\lambda_\theta \nonumber \\
1266: &\times&
1267: \frac{dt}{d\lambda}[r(\lambda_r),\theta(\lambda_\theta)]
1268: e^{i(k\omega_\theta + n\omega_r)\delta t(\lambda_r,\lambda_\theta)}\nonumber \\
1269: &\times&
1270: r(\lambda_r)
1271: e^{ik\Upsilon_\theta\lambda_\theta + in\Upsilon_r \lambda_r}~,\\
1272: {\tilde \theta_{kn}} &=&
1273: \frac{\Upsilon_r \Upsilon_\theta}{(2\pi)^2\Gamma}
1274: \int_0^{2\pi/\Upsilon_r} d\lambda_r
1275: \int_0^{2\pi/\Upsilon_\theta} d\lambda_\theta \nonumber \\
1276: &\times&
1277: \frac{dt}{d\lambda}[r(\lambda_r),\theta(\lambda_\theta)]
1278: e^{i(k\omega_\theta + n\omega_r)\delta t(\lambda_r,\lambda_\theta)}\nonumber \\
1279: &\times&
1280: \theta(\lambda_\theta)
1281: e^{ik\Upsilon_\theta\lambda_\theta + in\Upsilon_r \lambda_r}~,
1282: \end{eqnarray}
1283: where $r(\lambda_r)$ and $\theta(\lambda_\theta)$ are given
1284: by their Mino-time series expansions with $\lambda = \lambda_r$ and
1285: $\lambda = \lambda_\theta$, respectively,
1286: \begin{align}
1287: &r(\lambda_r) = \sum_n r_n e^{-i \Upsilon \lambda_r}~,&
1288: &\theta(\lambda_\theta) = \sum_k \theta_k e^{-i \Upsilon \lambda_\theta}~,&
1289: \end{align}
1290: and where
1291: \begin{equation}
1292: \delta t(\lambda_r,\lambda_\theta) =
1293: \sum_{kn} t_{kn} e^{-ik\Upsilon_\theta \lambda_\theta - in\Upsilon_r \lambda_r}~.
1294: \end{equation}
1295: The $\lambda$-derivative of $t$ can be evaluated analytically as given by
1296: Carter's first order geodesic equations, or it can be found from
1297: \begin{align}
1298: \frac{dt}{d\lambda}[r(\lambda_r),\theta(\lambda_\theta)] &= \Gamma& \nonumber \\
1299: &- i \sum_{kn}
1300: (k \Upsilon_\theta + n \Upsilon_r) t_{kn}
1301: e^{-ik\Upsilon_\theta \lambda_\theta - in\Upsilon_r \lambda_r}~.&
1302: \end{align}
1303:
1304: The derivation of the bi periodic form for $\phi(t)$ is slightly different. First,
1305: write $t(\lambda)$ as
1306: \begin{equation}
1307: t = \Gamma \lambda + \delta t~.
1308: \end{equation}
1309: Multiplying by $\omega_\phi$, and rearranging terms gives
1310: \begin{equation} \label{upl}
1311: \Upsilon_\phi \lambda = \Omega_\phi t - \Omega_\phi \delta t~,
1312: \end{equation}
1313: since $\omega_\phi \Gamma = \Upsilon_\phi$. Now write $\phi(\lambda)$ as
1314: \begin{equation}
1315: \phi = \Upsilon_\phi \lambda + \delta \phi~,
1316: \end{equation}
1317: and insert the above expression (\ref{upl}) for the first term to find
1318: \begin{align}
1319: \phi = \omega_\phi t + \widetilde{\delta\phi}~,
1320: \end{align}
1321: where
1322: \begin{equation}
1323: \widetilde{\delta\phi} = \delta \phi - \omega_\phi \delta t~.
1324: \end{equation}
1325: Treating $\widetilde{\delta\phi}$ as a function of $r$ and $\theta$ now
1326: allows it to be used in place of $f[r(\lambda), \theta(\lambda)]$ in
1327: Sec.~IV of \cite{drasco hughes 2004}. This gives the following
1328: expression for the coefficients of the bi periodic form of $\phi(t)$:
1329: \begin{eqnarray}
1330: {\tilde \phi_{kn}} &=&
1331: \frac{\Upsilon_r \Upsilon_\theta}{\Gamma (2\pi)^2}
1332: \int_0^{2\pi/\Upsilon_r} d\lambda_r
1333: \int_0^{2\pi/\Upsilon_\theta} d\lambda_\theta \nonumber \\
1334: &\times&
1335: \frac{dt}{d\lambda}[r(\lambda_r), \theta(\lambda_\theta)]
1336: e^{i(k\omega_\theta + n\omega_r)\delta t(\lambda_r,\lambda_\theta)}\nonumber \\
1337: &\times&
1338: \widetilde{\delta\phi}(\lambda_r,\lambda_\theta)
1339: e^{ik\Upsilon_\theta\lambda_\theta + in\Upsilon_r \lambda_r}~,
1340: \end{eqnarray}
1341: where
1342: \begin{equation}
1343: \widetilde{\delta\phi} (\lambda_r,\lambda_\theta) =
1344: \sum_{kn}(\phi_kn - \omega_\phi t_{kn})
1345: e^{-ik\Upsilon_\theta \lambda_\theta - in\Upsilon_r \lambda_r}~.
1346: \end{equation}
1347:
1348:
1349: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% BIBLIOGRAPHY %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1350:
1351: \begin{thebibliography}{0}
1352:
1353: %% Introduction
1354:
1355: \bibitem{will 2007}
1356: C. M. Will,
1357: %\emph{Testing the general relativistic ''no-hair'' theorems using the galactic center black hole SgrA*},
1358: Astrophys. J. Lett. {\bf 674}, L25 (2008).
1359:
1360: \bibitem{flanagan hughes 1998}
1361: {\'E}anna {\'E}. Flanagan, S. A. Hughes,
1362: %\emph{Measuring gravitational waves from binary black hole coalescences: I. Signal to noise for inspiral, merger, and ringdown},
1363: Phys. Rev. D {\bf 57}, 4535 (1998).
1364:
1365: \bibitem{dreyer et al}
1366: O. Dreyer \emph{et al},
1367: %\emph{Black Hole Spectroscopy: Testing General Relativity through Gravitational Wave Observations}
1368: Classical Quantum Gravity {\bf 21}, 787 (2004).
1369:
1370: \bibitem{berti et al 2006}
1371: E. Berti, V. Cardoso, C. M. Will,
1372: %\emph{On gravitational-wave spectroscopy of massive black holes with the space interferometer LISA},
1373: Phys. Rev. D {\bf 73}, 064030 (2006).
1374:
1375: \bibitem{berti et al 2007 ringdown}
1376: E. Berti, J. Cardoso, V. Cardoso, Marco Cavaglia,
1377: %\emph{Matched-filtering and parameter estimation of ringdown waveforms},
1378: arXiv:0707.1202v1 [gr-qc].
1379:
1380: \bibitem{gair et al 2004}
1381: J. R. Gair \emph{et al},
1382: %\emph{Event rate estimates for LISA extreme mass ratio capture sources},
1383: Classical Quantum Gravity {\bf 21}, S1595 (2004).
1384:
1385: \bibitem{hopman alexander 2006}
1386: C. Hopman and T. Alexander,
1387: %\emph{The effect of mass-segregation on gravitational wave sources near massive black holes},
1388: Astrophys. J {\bf 645}, L133 (2006).
1389:
1390: \bibitem{miller et al 2005}
1391: M. C. Miller, M. Freitag, D. P. Hamilton, and V. M. Lauburg,
1392: %\emph{Binary Encounters With Supermassive Black Holes: Zero-Eccentricity LISA Events},
1393: Astrophys. J {\bf 631}, L117 (2005).
1394:
1395: \bibitem{sigurdsson 2003}
1396: S. Sigurdsson,
1397: %\emph{Loss cone: past, present and future},
1398: Classical Quantum Gravity {\bf 20}, S45 (2003).
1399:
1400: \bibitem{barack cutler 2004}
1401: L. Barack and C. Cutler,
1402: %\emph{LISA Capture Sources: Approximate Waveforms, Signal-to-Noise Ratios, and Parameter Estimation Accuracy},
1403: Phys. Rev. D {\bf 69}, 082005 (2004).
1404:
1405: \bibitem{barack cutler 2007}
1406: L. Barack and C. Cutler
1407: %\emph{Using LISA EMRI sources to test off-Kerr deviations in the geometry of massive black holes},
1408: Phys. Rev. D {\bf 75}, 042003 (2007).
1409:
1410: \bibitem{gair mandel wen}
1411: J.\ R.\ Gair, I.\ Mandel, and L.\ Wen,
1412: %\emph{Improved time-frequency analysis of extreme-mass-ratio inspiral signals in mock LISA data},
1413: Class. Quant. Grav. {\bf 25}, 184031 (2008).
1414:
1415: \bibitem{gair babak porter barack}
1416: J.\ R.\ Gair, S.\ Babak, E.\ K.\ Porter, L.\ Barack,
1417: %\emph{A Constrained Metropolis-Hastings Search for EMRIs in the Mock LISA Data Challenge 1B},
1418: arXiv:0804.3322v1 [gr-qc].
1419:
1420: \bibitem{mldc}
1421: S. Babak \emph{et al},
1422: %\emph{Report on the second Mock LISA Data Challenge},
1423: Class. Quant. Grav. {\bf 25}, 114037 (2008).
1424:
1425: \bibitem{advanced ligo}
1426: Details on the design parameters for advanced LIGO can be found at
1427: http://www.ligo.caltech.edu/advLIGO~.
1428:
1429: \bibitem{brown et al 2006}
1430: D. A. Brown \emph{et al},
1431: %\emph{Gravitational waves from intermediate-mass-ratio inspirals for ground-based detectors},
1432: Phys. Rev. Lett. {\bf 99}, 201102 (2007).
1433:
1434: \bibitem{mandel et al 2007}
1435: I. Mandel, D. A. Brown, J. R. Gair, M. C. Miller,
1436: %\emph{Rates and Characteristics of Intermediate-Mass-Ratio Inspirals Detectable by Advanced LIGO},
1437: arXiv:0705.0285v1 [astro-ph].
1438:
1439: \bibitem{amaro-seoane et al}
1440: P. Amaro-Seoane \emph{et al},
1441: %\emph{Intermediate and extreme mass-ratio inspiralsÑastrophysics, science applications and detection using LISA},
1442: Class. Quant. Grav. {\bf 24}, R113 (2007).
1443:
1444: \bibitem{mino 2003}
1445: Y. Mino,
1446: %\emph{Perturbative approach to an orbital evolution around a supermassive black hole},
1447: Phys. Rev. D {\bf 67}, 084027 (2003).
1448:
1449: \bibitem{schmidt 2002}
1450: W. Schmidt,
1451: %\emph{Celestial mechanics in Kerr spacetime},
1452: Classical Quantum Gravity {\bf 19}, 2743 (2002).
1453:
1454: \bibitem{drasco hughes 2004}
1455: S. Drasco and S. A. Hughes,
1456: %\emph{Rotating black hole orbit functionals in the frequency domain},
1457: Phys. Rev. D 69, 044015 (2004).
1458:
1459: \bibitem{drasco 2006}
1460: S. Drasco,
1461: %\emph{Strategies for observing extreme mass ratio inspirals},
1462: Classical Quantum Gravity {\bf 23}, S769 (2006).
1463:
1464: \bibitem{hughes et al 2005}
1465: S. A. Hughes, S. Drasco, {\'E}. {\'E}. Flanagan, and J. Franklin,
1466: %\emph{Gravitational radiation reaction and inspiral waveforms in the adiabatic limit},
1467: Phys. Rev. Lett. {\bf 94}, 221101 (2005)
1468:
1469: \bibitem{drasco hughes 2006}
1470: S. Drasco and S. A. Hughes,
1471: %\emph{Gravitational wave snapshots of generic extreme mass ratio inspirals},
1472: Phys. Rev. D {\bf 73}, 024027 (2006).
1473:
1474: \bibitem{peters mathews}
1475: P. C. Peters and J. Mathews,
1476: %\emph{Gravitational Radiation from Point Masses in a Keplerian Orbit},
1477: Phys.\ Rev.\ {\bf 131}, 435 (1963).
1478:
1479: \bibitem{abramovici et al}
1480: A. Abramovici \emph{et al},
1481: %\emph{LIGO: The laser interferometer gravitational-wave observatory},
1482: Science {\bf 256}, 325 (1992).
1483:
1484: \bibitem{lisa science}
1485: For a review specific to LISA but generalizable to other detectors,
1486: see section 4 of the LISA science case document,
1487: \emph{LISA: Probing the Universe with Gravitational Waves},
1488: available as a \emph{mission document} at http://www.srl.caltech.edu/lisa/ .
1489:
1490: \bibitem{hughes 2006}
1491: S. A. Hughes,
1492: %\emph{(Sort of) Testing relativity with extreme mass ratio inspirals},
1493: AIP Conf.~Proc.~{\bf 873}, 233 (2006).
1494: arXiv:gr-qc/0608140v1.
1495:
1496: \bibitem{psaltis et al 2007}
1497: D. Psaltis, D. Perrodin, K. R. Dienes, and I. Mocioiu,
1498: %\emph{Kerr Black Holes are Not Unique to General Relativity},
1499: Phys. Rev. Lett. {\bf 100}, 091101 (2008).
1500:
1501: \bibitem{li lovelace 2007}
1502: C. Li and G. Lovelace,
1503: %\emph{A generalization of Ryan's theorem: probing tidal coupling with gravitational waves from nearly circular, nearly equatorial, extreme-mass-ratio inspirals},
1504: Phys. Rev. D {\bf 77}, 064022 (2008).
1505:
1506: \bibitem{ryan}
1507: F. D. Ryan,
1508: %\emph{Gravitational waves from the inspiral of a compact object into a massive, axisymmetric body with arbitrary multipole moments},
1509: Phys. Rev. D {\bf 52}, 5707 (1995).
1510:
1511: %% EMRI-Snapshot spectra
1512:
1513: \bibitem{drasco flanagan hughes 2005}
1514: S. Drasco, {\' E}. {\' E}. Flanagan, and S. A. Hughes,
1515: %\emph{Computing inspirals in Kerr in the adiabatic regime. I. The scalar case},
1516: Classical Quantum Gravity {\bf 22}, S801 (2005).
1517:
1518: \bibitem{mtw}
1519: C.\ W.\ Misner, K.\ S.\ Thorne, and J.\ A.\ Wheeler,
1520: \emph{Gravitation} (Freeman, San Francisco, 1973).
1521:
1522: \bibitem{carter}
1523: B.\ Carter,
1524: %\emph{Global structure of the Kerr family of gravitational fields},
1525: Phys.\ Rev.\ {\bf 174}, 1559 (1968).
1526:
1527: \bibitem{pound poisson nickel 2005}
1528: A. Pound, E. Poisson, and B. G. Nickel,
1529: %\emph{Limitations of the adiabatic approximation to the gravitational self-force},
1530: Phys. Rev. D {\bf 72}, 124001 (2005).
1531:
1532: \bibitem{pound poisson 2007a}
1533: A. Pound and E. Poisson,
1534: %\emph{Multi-scale analysis of the electromagnetic self-force in a weak gravitational field},
1535: Phys. Rev. D {\bf 77}, 044012 (2008).
1536:
1537: \bibitem{pound poisson 2007b}
1538: A. Pound and E. Poisson,
1539: %\emph{Osculating orbits in Schwarzschild spacetime, with an application to extreme mass-ratio inspirals},
1540: Phys. Rev. D {\bf 77}, 044013 (2008).
1541:
1542: \bibitem{teukolsky}
1543: S. A. Teukolsky,
1544: %\emph{Rotating Black Holes: Separable Wave Equations for Gravitational and Electromagnetic Perturbations},
1545: Phys. Rev. Lett. {\bf 29}, 1114 (1972);
1546: S. A. Teukolsky,
1547: %\emph{Perturbations of a rotating black hole. I. fundamental equations for gravitational, electromagnetic, and neutrino-field perturbations},
1548: Astrophys. J. {\bf 185}, 635 (1973);
1549:
1550: \bibitem{andersson frolov novikov}
1551: N. Andersson, V. P. Frolov, and I. D. Novikov, in
1552: \emph{Black hole physics, basic concepts and new developments}
1553: (Kluwer Academic Publishers, Dordrecht, 1998).
1554:
1555: \bibitem{poisson sasaki}
1556: E. Poisson and M. Sasaki,
1557: %\emph{Gravitational radiation from a particle in circular orbit around a black hole. V. Black-hole absorption and tail corrections},
1558: Phys. Rev. D {\bf 51}, 5753 (1995).
1559:
1560: \bibitem{sago et al 2005}
1561: N. Sago, T. Tanaka, W. Hikida, and H. Nakano,
1562: %\emph{Adiabatic radiation reaction to the orbits in Kerr Spacetime},
1563: Prog. Theor. Phys. {\bf 114}, 509 (2005).
1564:
1565: \bibitem{sago et al 2006}
1566: N. Sago \emph{et al},
1567: %\emph{The adiabatic evolution of orbital parameters in the Kerr spacetime},
1568: Prog. Theor. Phys. {\bf 115}, 873 (2006).
1569:
1570: %% Sample spectra
1571:
1572: \bibitem{mino 2007}
1573: Y. Mino,
1574: %\emph{Modulation of the gravitational waveform by the effect of radiation reaction},
1575: Phys. Rev. D {\bf 77}, 044008 (2008).
1576:
1577: \bibitem{van den broeck and sengupta}
1578: C. Van Den Broeck and A. S. Sengupta,
1579: %\emph{Binary black hole spectroscopy},
1580: Classical Quantum Gravity {\bf 24}, 1089 (2007).
1581:
1582: \bibitem{sintes vecchio}
1583: A. M. Sintes and A. Vecchio,
1584: %\emph{LISA observations of massive black holes binaries using post-Newtonian wave-forms},
1585: arXiv:gr-qc/0005059v1.
1586:
1587: \bibitem{hellings moore}
1588: R. W. Hellings and T. A. Moore,
1589: %\emph{The information content of gravitational wave harmonics in compact binary inspiral},
1590: Classical Quantum Gravity {\bf 20}, S181 (2003).
1591:
1592: \bibitem{arun et al}
1593: K. G. Arun, B. R. Iyer, B. S. Sathyaprakash, and S. Sinha,
1594: %\emph{Higher harmonics increase LISA's mass reach for supermassive black holes},
1595: Phys. Rev. D {\bf 75}, 124002 (2007).
1596:
1597: \bibitem{arun et al 2}
1598: K. G. Arun \emph{et al}
1599: %\emph{Higher signal harmonics, LISA's angular resolution, and dark energy},
1600: Phys. Rev. D {\bf 76}, 104016 (2007).
1601:
1602: \bibitem{trias sintes}
1603: M. Trias and A. M. Sintes,
1604: %\emph{LISA observations of supermassive black holes: parameter estimation using full post-Newtonian inspiral waveforms},
1605: Phys. Ref. D {\bf 77}, 024030 (2008).
1606:
1607: \bibitem{berti et al 2007}
1608: E. Berti \emph{et al},
1609: %\emph{Inspiral, merger and ringdown of unequal mass black hole binaries: a multipolar analysis},
1610: Phys. Rev. D {\bf 76}, 064034 (2007).
1611:
1612: \bibitem{vaishnav et al}
1613: B. Vaishnav, I. Hinder, F. Herrmann, and D. Shoemaker,
1614: %\emph{Matched Filtering of Numerical Relativity Templates of Spinning Binary Black Holes},
1615: Phys. Rev. D {\bf 76}, 084020 (2007).
1616:
1617: \bibitem{cutler}
1618: C. Cutler,
1619: %\emph{Angular resolution of the LISA gravitational wave detector},
1620: Phys. Rev. D {\bf 57}, 7089 (1998).
1621:
1622: \bibitem{vecchio}
1623: A. Vecchio,
1624: %\emph{LISA observations of rapidly spinning massive black hole binary systems},
1625: Phys. Rev. D {\bf 70}, 042001 (2004).
1626:
1627: \bibitem{lang hughes}
1628: R. N. Lang and S. A. Hughes,
1629: %\emph{Localizing coalescing massive black hole binaries with gravitational waves},
1630: Astrophys. J. {\bf 677}, 1184 (2008).
1631:
1632: \bibitem{berti buonanno will}
1633: E. Berti, A. Buonanno, and C. M. Will,
1634: %\emph{Estimating spinning binary parameters and testing alternative theories of gravity with LISA},
1635: Phys. Rev. D {\bf 71}, 084025 (2005).
1636:
1637:
1638: %% Spectra from an kludged inspiral
1639:
1640: \bibitem{gair glampedakis 2005}
1641: J. R. Gair and K. Glampedakis,
1642: %\emph{Improved approximate inspirals of test-bodies into Kerr black holes},
1643: Phys. Rev. D {\bf 73}, 064037 (2006).
1644:
1645: \bibitem{babak et al 2007}
1646: %S. Babak \emph{et al},
1647: S. Babak, H. Fang, J. R. Gair, K. Glampedakis, and S. A. Hughes
1648: %\emph{``Kludge'' gravitational waveforms for a test-body orbiting a Kerr black hole},
1649: Phys. Rev. D {\bf 75}, 024005 (2007).
1650:
1651: %% Verification of black hole orbits
1652:
1653: \bibitem{gair li mandel 2007}
1654: J. R. Gair, C. Li, and I. Mandel,
1655: %\emph{Observable Properties of Orbits in Exact Bumpy Spacetimes},
1656: Phys. Rev. D {\bf 77}, 024035 (2008).
1657:
1658: \bibitem{flanagan hinderer 2007}
1659: {\' E}. {\' E}. Flanagan and T. Hinderer,
1660: %\emph{Evolution of the Carter constant for inspirals into a black hole: effect of the black hole quadrupole},
1661: Phys. Rev. D {\bf 75}, 124007 (2007).
1662:
1663: \bibitem{gair et al 2007}
1664: J. R. Gair, I. Mandel, and L. Wen,
1665: %\emph{Time-frequency analysis of extreme-mass-ratio inspiral signals in mock LISA data},
1666: arXiv:0710.5250v1 [gr-qc].
1667:
1668: %% Conclusion
1669:
1670: \bibitem{moreno-garrido mediavilla buitrago}
1671: C. Moreno-Garrido, E. Mediavilla, and J. Buitrago,
1672: %\emph{Gravitational radiation from point masses in elliptical orbits: spectral analysis and orbital parameters},
1673: Mon. Not. R. Astron. Soc. {\bf 274}, 115 (1995).
1674:
1675: \bibitem{poisson}
1676: E. Poisson,
1677: %\emph{Gravitational radiation from a particle in circular orbit around a black hole. I. Analytical results for the nonrotating case},
1678: Phys. Rev. D {\bf 47}, 1497 (1993).
1679:
1680: \bibitem{ganz et al}
1681: K. Ganz, W. Hikida, H. Nakano, N. Sago, and T. Tanaka,
1682: %\emph{Adiabatic Evolution of three 'Constants' of Motion for Greatly Inclined Orbits in Kerr spacetime},
1683: Prog. Theor. Phys. {\bf 117}, 1041 (2007).
1684:
1685: \bibitem{miller}
1686: J. M. Miller,
1687: %\emph{Relativistic X-ray Lines from the Inner Accretion Disks Around Black Holes},
1688: ARA{\&}A {\bf 45}, 441 (2007).
1689:
1690: \bibitem{reis et al}
1691: R. C. Reis \emph{et al},
1692: %\emph{A systematic look at the very high and low/hard state of GX 339?4: constraining the black hole spin with a new reflection model},
1693: Mon. Not. R. Astron. Soc. {\bf 387}, 1489 (2008).
1694:
1695: \end{thebibliography}
1696:
1697: \end{document}
1698: