astro-ph0503672/ms.tex
1: 
2: 
3: \documentclass[english]{emulateapj}
4: %\documentclass[12pt,preprint]{aastex}
5: 
6: \usepackage{times}
7: \slugcomment{ApJ accepted}
8: 
9: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%% LyX specific LaTeX commands.
10: %% Because html converters don't know tabularnewline
11: \providecommand{\tabularnewline}{\\}
12: 
13: \newcommand{\peryr}{{\rm yr^{-1}}}
14: \newcommand{\yr}{{\rm yr}}
15: \newcommand{\pc}{\mathrm{pc}}
16: \newcommand{\perpc}{\mathrm{pc}^{-1}}
17: 
18: \newcommand{\LISA}{{\it LISA}}
19: 
20: \newcommand{\Mbh}{M_{\bullet}}
21: 
22: \newcommand{\Mo}{M_{\odot}}
23: 
24: \newcommand{\Ro}{R_{\odot}}
25: 
26: \newcommand{\Lo}{L_{\odot}}
27: 
28: \newcommand{\Ms}{M_{\star}}
29: 
30: \newcommand{\Rs}{R_{\star}}
31: 
32: \newcommand{\Ls}{L_{\star}}
33: 
34: \newcommand{\Es}{E_{\star}}
35: 
36: \newcommand{\ns}{N_{\bullet}}
37: 
38: \newcommand{\scr}{s_{\mathrm{crit}}}
39: 
40: \newcommand{\nbh}{n_{\bullet}}
41: 
42: \newcommand{\mF}{\mathcal{F}}
43: 
44: 
45: 
46: \shorttitle{INSPIRAL AND RELAXATION NEAR A MBH}
47: \shortauthors{HOPMAN \& ALEXANDER}
48:  
49: \bibpunct{}{}{}{a}{}{}
50: 
51: \begin{document}
52: 
53: \title{The orbital statistics of stellar inspiral and relaxation near a
54: massive black hole: characterizing gravitational wave sources}
55: 
56: 
57: \author{Clovis Hopman\altaffilmark{1} and Tal Alexander\altaffilmark{1,2}}
58: 
59: 
60: \altaffiltext{1}{Center for Astrophysics, Faculty of Physics, Weizmann
61: Institute of Science, POB 26, Rehovot 76100, Israel;
62: clovis.hopman@weizmann.ac.il, tal.alexander@weizmann.ac.il}
63: \altaffiltext{2}{Incumbent of the William Z. \& Eda Bess Novick career
64: development chair}
65: 
66: 
67: 
68: \begin{abstract}
69: We study the orbital parameters distribution of stars that are
70: scattered into nearly radial orbits and then spiral into a massive
71: black hole (MBH) due to dissipation, in particular by emission of
72: gravitational waves (GW). This is important for GW detection, e.g. by
73: the \emph{Laser Interferometer Space Antenna} (\emph{\LISA}). Signal
74: identification requires knowledge of the waveforms, which depend on
75: the orbital parameters. We use analytical and Monte Carlo methods to
76: analyze the interplay between GW dissipation and scattering in the
77: presence of a mass sink during the transition from the initial
78: scattering-dominated phase to the final dissipation-dominated phase of
79: the inspiral. Our main results are (1) Stars typically enter the
80: GW-emitting phase with high eccentricities. (2) The GW event rate per
81: galaxy is $\mathrm{few\!\times\!10^{-9}}\,\mathrm{yr^{-1}}$ for
82: typical central stellar cusps, almost independently of the relaxation
83: time or the MBH mass. (3) For intermediate mass black holes (IBHs) of
84: $\sim\!10^{3}\,\Mo$ such as may exist in dense stellar clusters, the
85: orbits are very eccentric and the inspiral is rapid, so the sources
86: are very short-lived.
87: \end{abstract}
88: 
89: \keywords{black hole physics --- stellar dynamics --- gravitational waves}
90: 
91: 
92: \section{Introduction}\label{s:intro}
93: 
94: \setcounter{footnote}{0}
95: 
96: Dissipative interactions between stars and massive black holes (MBHs;
97: $\Mbh\!\gtrsim\!10^{6}\,\Mo$) in galactic nuclei (e.g. Gebhardt et
98: al. \cite{Geb00}, \cite{Geb03}), or intermediate mass black holes
99: (IBHs; $10^{2}\!<\!\Mbh\!\lesssim\!10^{4}\,\Mo$), which may exist
100: in dense stellar clusters, have been in the focus of several recent
101: studies.  The interest is mainly motivated by the possibility of
102: using the dissipated power to detect the BH, or to probe General Relativity.
103: Examples of such processes are tidal heating (e.g. Alexander \& Morris
104: \cite{AM03}; Hopman, Portegies Zwart \& Alexander \cite{HPZA04})
105: and gravitational wave (GW) radiation (Hils \& Bender \cite{HB95};
106: Sigurdsson \& Rees \cite{SR97}; Ivanov \cite{IV02}; Freitag \cite{FR01},
107: \cite{FR03}).
108: 
109: A statistical characterization of inspiral orbits is of interest in
110: anticipation of GW observations by the Laser Interferometer Space
111: Antenna (\LISA). \textit{\LISA\ } will be able to observe GW from
112: stars at cosmological distances during the final, highly relativistic
113: phase of inspiral into a $\sim\!10^{6}\,\Mo$ MBH, thereby opening
114: a new non-electromagnetic astronomical window. GW from inspiraling
115: compact objects (COs) is one of the three major targets of the \LISA\ 
116: mission (Barack \& Cutler \cite{BC04a}, \cite{BC04b}; Gair et al.
117: \cite{Gai04}), together with cosmological MBH--MBH mergers and Galactic
118: CO--CO mergers.
119: 
120: \LISA\ can detect GW emission from stars with orbital
121: period shorter than $P_{L}\!\sim\!10^{4}\,\mathrm{s}$.  In order for
122: the shortest possible period to be small enough to be detectable by
123: \LISA\ , the MBH has to be of moderate mass,
124: $\Mbh\!\lesssim\!5\!\times\!10^{6}\,\Mo$ (Sigurdsson \& Rees
125: \cite{SR97}). \emph{\LISA} is expected to be able to detect inspiral
126: into MBHs of $10^{6}\,\Mo$ to distances as far as $\gtrsim$ 1 Gpc.
127: 
128: The detailed time-evolution of the GW depends on the eccentricity of
129: the stellar orbit, and therefore probes both General Relativity and
130: the statistical predictions of stellar dynamics theory. Due to the low
131: signal to noise ratio, knowledge of the wave forms is required in
132: advance. For this purpose, it is necessary to estimate the orbital
133: characteristics of the GW-emitting stars, and in particular the
134: distribution function (DF) of their eccentricities (Pierro et
135: al. \cite{PPSLR01}; Glampedakis, Hughes \& Kennefick \cite{GHK02}), as
136: the wave forms are strong functions of the eccentricity (e.g. Barack
137: \& Cutler \cite{BC04a}; Wen \& Gair \cite{WG05}).  This study focuses
138: on inspiral by GW emission. However, it should be emphasized that
139: inspiral is a general consequence of dissipation, and the formalism
140: presented below can be extended in a straight-forward way to other
141: dissipation processes, such as tidal heating.
142: 
143: The \emph{prompt infall} of a star into a MBH and its destruction
144: have been studied extensively (\S\ref{ss:prompt}). Here we analyze
145: a different process, the \emph{slow inspiral} of stars (Alexander
146: \& Hopman \cite{AH03}). A star on a highly eccentric orbit with small
147: periapse $r_{p}$, repeatedly loses some energy $\Delta E$ every
148: periapse passage due to GW emission, and its orbit gradually decays.
149: At a distance $r_{0}$ from the MBH, where the orbital period is $P_{0}$,
150: the time-scale $t_{0}$ for completing the inspiral (i.e. decaying
151: to a $P\!\rightarrow\!0$ orbit) is much longer than the time-scale
152: $P_{0}$ needed to reach the MBH directly on a nearly radial orbit.
153: While the orbit decays, two-body scatterings by other stars continually
154: perturb it, changing its orbital angular momentum $J$ by order unity
155: on a timescale $t_{J}$. Because $t_{0}\!\gg\! P_{0}$, inspiraling
156: stars are much more susceptible to scattering than those on infall
157: orbits. If $t_{0}\!>\! t_{J}$, either because $r_{0}$ is large or
158: $r_{p}$ is large (small $\Delta E$), then the orbit will not have
159: time to decay and reach an observationally interesting short period.
160: Before that can happen, the star will either be scattered to a wider
161: orbit where energy dissipation is no longer efficient, or conversely,
162: plunge into the MBH. Inspiral is thus much rarer than direct infall.
163: The stellar consumption rate, and hence the properties of the stellar
164: distribution function (DF) at low $J$, are dominated by prompt infall,
165: with inspiral contributing only a small correction. This DF describes
166: the parent population of the inspiraling stars.
167: 
168: We show below that the DF of the small subset of stars on low-$J$
169: orbits that complete the inspiral and are GW sources is very different
170: from that of the parent population (Fig. \ref{f:Jlossdist}). This
171: results from the interplay between GW dissipation and scattering in
172: the presence of a mass sink during the transition ($t_{0}\!\sim\! t_{J}$)
173: from the initial scattering-dominated phase to the final dissipation-dominated
174: phase of the inspiral.
175: 
176: This paper is organized as follows. In \S\ref{s:lc} we recapitulate
177: some of the results of loss cone theory for the prompt infall, and
178: extend it to slow inspiral. In \S\ref{s:inspscat} and \S\ref{s:approaches}
179: we present a detailed analytical discussion of the main effects that
180: determine the rates of GW events and their statistical properties.
181: In \S\ref{s:approaches} we describe three different approaches for
182: studying the problem: by Monte Carlo simulations, by solving numerically
183: the 2D diffusion / dissipation equation in $E$ and $J$, and by a
184: simplified analytical model, which mimics the behavior of a typical
185: star. In \S\ref{s:results} we apply the MC simulation to a MBH in
186: a galactic nucleus, and an IBH in a stellar cluster. We summarize
187: our results in \S\ref{s:sum}.
188: 
189: 
190: \section{The loss-cone}
191: 
192: \label{s:lc}
193: 
194: The rate at which stars are consumed by an MBH and the effect this
195: has on the stellar DF near it have been studied extensively (Peebles
196: \cite{P72}; Frank \& Rees \cite{FR76}; Bahcall \& Wolf \cite{BW76},
197: \cite{BW77}; Lightman and Shapiro \cite{LS77}; Cohn \& Kulsrud \cite{CK78};
198: Syer \& Ulmer \cite{SU99}; Magorrian \& Tremaine \cite{MT99}; Miralda-Escud\'{e}
199: \& Gould \cite{MG00}; Freitag \& Benz \cite{FB01}; Alexander \& Hopman
200: \cite{AH03}; Wang \& Merritt \cite{WM04}; see Sigurdsson \cite{S03}
201: for a comparative review).  Self-consistent N-body simulations with
202: stellar captures were recently performed by Baumgardt, Makino, \&
203: Ebisuzaki (\cite{Baum04a}, \cite{Baum04b}) and by Preto, Merritt
204: \& Spurzem (\cite{P04}).
205: 
206: We begin by summarizing these results, neglecting dissipative processes.
207: We then extend the formalism to include dissipative processes.
208: 
209: 
210: \subsection{Prompt infall}
211: 
212: \label{ss:prompt}
213: 
214: The stellar orbits are defined by a specific angular momentum $J$ and
215: relative specific energy $\varepsilon\!=\!\psi(r)\!-\! v^{2}/2$
216: (hereafter {}``angular momentum'' and {}``energy''), where $\psi$ is
217: the relative gravitational potential, and $v$ is the velocity of a
218: star with respect to the MBH. A spherical mass distribution and a
219: nearly spherical velocity distribution are assumed.
220: 
221: 
222: 
223: Orbits in the Schwarzschild metric (unlike Keplerian orbits) can escape
224: the MBH only if their angular momentum is high enough, $J\!>\! J_{lc}(\varepsilon)$.
225: The phase space volume $J\!<\! J_{lc}$ is known as the {}``loss-cone''.
226: As argued below, stars that are scattered to low-$J$ orbits are typically
227: on nearly zero-energy orbits. For such orbits
228: 
229: \begin{equation}
230: J_{lc}(\varepsilon\!=\!0)=\frac{4G\Mbh}{c}\,,\label{e:Jlc}\end{equation}
231: The size of the loss-cone $J_{lc}$ is nearly constant over the
232: relevant range of $\varepsilon$.  Only during the very last in-spiral
233: phase the energy of the star becomes non-negligible compared to its
234: rest-mass, in which case the loss-cone is slightly modified (see
235: section {[}\ref{sub:MC}{]}). Deviations from geodetic motion due to
236: tidal interactions are neglected here.  This assumption is justified
237: for COs orbiting $\sim\!10^{6}\,\Mo$ MBHs, where the tidal radius is
238: much smaller than the event horizon. For main-sequence (MS) stars
239: where the tidal radius lies outside the event horizon, the loss-cone
240: is similarly defined as the minimal $J$ required to avoid tidal
241: disruption.
242: 
243: Stars that are initially on orbits with $J\!<\! J_{lc}$ will promptly
244: fall into the MBH on an orbital timescale. Subsequently, the infall
245: flow in $J$-space, $\mF(\varepsilon;J)$, is set by the rate at which
246: relaxation processes (here assumed to be multiple two-body scattering
247: events) re-populate the loss-cone orbits. 
248: 
249: Diffusion in $\varepsilon$-space occurs on the relaxation timescale,
250: $t_{r}\!\sim\!\varepsilon/\dot{\varepsilon}$, whereas diffusion in
251: $J$-space occurs on the angular momentum relaxation timescale,
252: \begin{equation} t_{J}\sim
253: J^{2}/\dot{\left(J^{2}\right)}\sim[J/J_{m}(\varepsilon)]^{2}t_{r}\,,\label{e:tJ}\end{equation}
254: where $J_{m}(\varepsilon)$ is the maximal (circular orbit) angular
255: momentum for specific energy $\varepsilon$.The square root dependence
256: of $J$ on $t_{J}$ reflects the random walk nature of the process.
257: Typically, $J_{lc}\!\ll\! J_{m}$. In principle, stars can enter the
258: loss-cone, $J\!<\! J_{lc}(\varepsilon)$ either by a decrease in $J$,
259: or by an increase in $\varepsilon$ (up to the last stable orbit).  In
260: practice, diffusion in $J$-space is much more efficient: the energy of
261: a star must increase by many orders of magnitude in order for it to
262: reach the loss-cone, which takes many relaxation times. The angular
263: momentum of the star, on the other hand, needs only to change by order
264: unity in order for the star to be captured, which happens on a much
265: shorter time $t_{J}\leq t_{r}$.
266: 
267: The ratio between $J_{lc}$ and the mean change in angular momentum
268: per orbit, $\Delta J$, defines two dynamical regimes of loss-cone
269: re-population (Lightman \& Shapiro \cite{LS77}). In the {}``Diffusive
270: regime'' of stars with large $\varepsilon$ (tight orbits), $\Delta J\!\ll\! J_{lc}$
271: and so the stars slowly diffuse in $J$-space. The loss-cone remains
272: nearly empty at all times since any star inside it is promptly swallowed.
273: At $J\!\gg\! J_{lc}$ the DF is nearly isotropic, but it falls logarithmically
274: to zero at $J\gtrsim J_{lc}$ (Eq. \ref{e:Nlc}). In the {}``full
275: loss-cone regime'' (sometimes also called the {}``pinhole'' or
276: {}``kick'' regime) of stars with small $\varepsilon$ (wide orbits),
277: $\Delta J\!\gg\! J_{lc}$ and so the stars can enter and exit the
278: loss-cone many times before reaching periapse. As a result, the DF
279: is nearly isotropic at all $J\!\gtrsim\! J_{lc}$. We argue below
280: that only the diffusive regime is relevant for inspiral.
281: 
282: The DF in the diffusive regime is described by the Fokker-Planck equation.
283: We follow Lightman \& Shapiro (\cite{LS77}), who neglect the small
284: contribution of energy diffusion to $\mF$, and write the Fokker-Planck
285: equation for the number density of stars $N(\varepsilon,J;t)$ as%
286: \footnote{We use the notation $N(x;y)$, where $x$ stands for any argument
287: or set of arguments (scalar of vector) and $y$ is a parameter, to
288: denote the stellar number density per $\mathrm{d}x$ interval at $y$.
289: The units of $N$ are the inverse of those of $x$. For example, $N(\varepsilon;t)\!=\!\int_{0}^{J_{m}}\mathrm{d}JN(\varepsilon,J;t)$
290: is the stellar number density per unit specific energy at time $t$.%
291: }
292: 
293: \begin{equation}
294: \frac{\partial N(\varepsilon,J;t)}{\partial
295:  t}=-\frac{\partial\mF(\varepsilon;J)}{\partial
296:  J},\label{e:FP}\end{equation} where \begin{equation}
297:  \mF(\varepsilon;J)=N(\varepsilon,J)\langle\Delta
298:  J\rangle-\frac{1}{2}\frac{\partial}{\partial
299:  J}N(\varepsilon,J)\langle\Delta J^{2}\rangle.\end{equation} The
300:  diffusion coefficients $\langle\Delta J\rangle$ and $\langle\Delta
301:  J^{2}\rangle$ obey the relation \begin{equation} \langle\Delta
302:  J^{2}\rangle=2J\langle\Delta J\rangle\qquad(J\!\ll\!
303:  J_{m})\,,\label{eq:Jcoeffs}\end{equation} (Lightman \& Shapiro
304:  \cite{LS77}; Magorrian \& Tremaine \cite{MT99}).  Much of the
305:  difficulty in obtaining an exact solution for the Fokker-Planck
306:  equation stems from the dependence of the diffusion coefficients on
307:  the DF; self-consistency requires solving a set of coupled equations.
308:  For many practical applications the diffusion coefficients are
309:  estimated in a non-self-consistent way, for example by assuming local
310:  homogeneity and isotropy (e.g. Binney \& Tremaine
311:  \cite{BT87}). Irrespective of its exact form, $\langle\Delta
312:  J^{2}\rangle$ describes a random-walk process, and is therefore
313:  closely related to the relaxation time.  In anticipation of the
314:  eventual necessity of introducing such approximations, we forgo from
315:  the outset the attempt to write down explicit expressions for the
316:  diffusion coefficients. Instead, we use them to \emph{define} the
317:  relaxation time $t_{r}$ as the time required to diffuse in $J^{2}$ by
318:  $J_{m}^{2}$,
319: 
320: \begin{equation}
321: t_{r}(\varepsilon)=\frac{J_{m}^{2}(\varepsilon)}{\langle\Delta
322: J^{2}\rangle}\,.\label{e:diffcoeff}\end{equation} We then treat the
323: relaxation time as a free parameter that characterizes the system's
324: typical timescale for the evolution of the DF, and whose value can be
325: estimated by Eqs. (\ref{e:tr_rh}, \ref{e:tr}) below.  For simplicity,
326: the relaxation time $t_{r}$ is assumed to be independent of angular
327: momentum%
328: \footnote{In general, if the relaxation time depends on $r$, this will
329: introduce some dependence of $t_r$ on $J$. We do not consider this
330: dependence here.}, but can generally be a function of energy.
331: 
332: 
333: At steady state, the stellar current $\mF(\varepsilon)$ is independent
334: of $J$,
335: 
336: \begin{equation}
337: \mF(\varepsilon)=\frac{1}{2}\frac{J_{m}^{2}(\varepsilon)}{t_{r}}\left[\frac{N(\varepsilon,J)}{J}-\frac{\partial N(\varepsilon,J)}{\partial J}\right]\,.\label{e:Fss}\end{equation}
338: Solving this equation yields\begin{equation}
339: N(\varepsilon,J)=-\frac{2\mF t_{r}}{J_{m}(\varepsilon)^{2}}J\ln J+CJ\,.\label{e:solnoBC}\end{equation}
340:  The integration constants $C$ and $\mF$ that are determined by
341: the boundary conditions $N(\varepsilon,J_{lc})\!=\!0$ and $N(\varepsilon,J_{m})\!=\! N_{\mathrm{iso}}(\varepsilon,J_{m})$.
342: The isotropic DF is separable in $\varepsilon$ and $J$,
343: 
344: \begin{equation}
345: N_{\mathrm{iso}}(\varepsilon,J)\mathrm{d}\varepsilon\mathrm{d}J=\frac{2N_{\mathrm{iso}}(\varepsilon)J}{J_{m}^{2}(\varepsilon)}\mathrm{d}\varepsilon\mathrm{d}J\,.\label{e:Niso}\end{equation}
346:  Applying these boundary conditions\footnote{Adiabatic MBH growth may
347:  lead to some anisotropy (Quinlan, Hernquist \& Sigurdsson
348:  \cite{QHS95}.). Here we assume that far from the loss-cone the DF will
349:  be isotropic} to equation (\ref{e:solnoBC}), the DF is given by
350: 
351: \begin{equation}
352: N(\varepsilon,J)=\frac{2N_{\mathrm{iso}}(\varepsilon)J}{J_{m}^{2}(\varepsilon)}\frac{\ln(J/J_{lc})}{\ln(J_{m}/J_{lc})},\label{e:Nlc}\end{equation}
353:  and the stellar current into the MBH per energy interval is
354: 
355: \begin{equation}
356: \mF(\varepsilon)=\frac{N_{\mathrm{iso}}(\varepsilon)}{\ln(J_{m}/J_{lc})t_{r}(\varepsilon)}\,.\label{e:Flow}\end{equation}
357:  Note that the capture rate in the diffusive regime depends only logarithmically
358: on the size of the loss cone. 
359: 
360: The prompt infall rate $\Gamma_{p}$ in the diffusive regime is then
361: given by
362: 
363: \begin{equation}
364: \Gamma_{p}=\int_{\varepsilon_{p}}^{\infty}\frac{\mathrm{d}\varepsilon N_{\mathrm{iso}}(\varepsilon)}{\ln(J_{m}/J_{lc})t_{r}(\varepsilon)}\,,\label{e:Gp}\end{equation}
365:  where the energy $\varepsilon_{p}$ separates the diffusive and full
366: loss-cone regimes. A star samples all angular momenta $J_{lc}\!<\! J\!<\! J_{m}$
367: in a relaxation time, and it is promptly captured once $J\!<\! J_{lc}$.
368: The total rate is therefore of order $\Gamma_{p}\!\sim\! N_{\mathrm{iso}}(<\! a_{p})/t_{r}(a_{p})$,
369: where $a_{p}$ is the typical radius associated with orbits of energy
370: $\varepsilon_{p}$ ($a_{p}$ is the semi-major axis for Keplerian
371: orbits, see \S\ref{ss:KeplerCusp}) and $N_{\mathrm{iso}}(<\! a_{p})$
372: is the number of stars within $a_{p}$. The rate is logarithmically
373: suppressed because of the diluted occupation of phase space near the
374: loss cone.
375: 
376: 
377: \subsection{Keplerian orbits in a power-law cusp}
378: 
379: \label{ss:KeplerCusp}
380: 
381: The MBH dominates the stellar potential within the radius of influence,
382: 
383: \begin{equation}
384: r_{h}=\frac{G\Mbh}{\sigma^{2}},\label{e:rh}\end{equation} where
385:  $\sigma^{2}$ is the 1D stellar velocity dispersion far from the
386:  MBH. The mass enclosed within $r_{h}$ is roughly equal to $\Mbh$.
387:  Various formation scenarios predict that the spatial stellar number
388:  density at $r\!<\! r_{h}$ should be approximately a power law (e.g.
389:  Bahcall \& Wolf \cite{BW76}; Young \cite{You80}) \begin{equation}
390:  n_{\star}(r)=\frac{(3/2-p)N_{h}}{4\pi
391:  r_{h}^{3}}\left(\frac{r}{r_{h}}\right)^{-3/2-p}\,,\label{e:nplaw}\end{equation}
392:  where $N_{h}$ is the number of stars inside $r_{h}$. This corresponds
393:  to an energy distribution
394:  $N(\varepsilon)\mathrm{d}\varepsilon\!\propto\!\varepsilon^{p-5/2}\mathrm{d}\varepsilon$.
395:  A stellar cusp with $p\!\sim\!0$ has been observed in the Galactic
396:  Center (Alexander \cite{TA99}; Genzel et al. \cite{Genzelea03}).  For
397:  a single mass population, it was shown by analytical considerations
398:  that $p=1/4$ (Bahcall \& Wolf \cite{BW76}). This has been confirmed
399:  recently by N-body simulations (Baumgardt et al. \cite{Baum04a};
400:  Preto et al. \cite{P04}).
401: 
402: Mass segregation drives the heavy stars in the population to the center.
403: The radial distribution of the different mass components can be then
404: approximated as average power-laws, steeper ($p\!>\!0$) for the heavier
405: masses and flatter ($p\!\lesssim\!0$) for the lower masses (Bahcall \&
406: Wolf \cite{BW77}; Baumgardt et al. \cite{Baum04b}).
407: 
408: Typically, the diffusive regime is within the radius of influence.
409: We therefore assume from this point on that the stars move on Keplerian
410: orbits ($\psi\!=\! G\Mbh/r$) in a power-law density cusp. The stellar
411: orbits are characterized by a semi-major axis $a$, eccentricity $e$,
412: periapse $r_{p}$ and period $P$,
413: 
414: \begin{eqnarray}
415: a\!=\frac{G\Mbh}{2\varepsilon}\,,\qquad\:\:\:\:\:\: &  & e^{2}\!=\!1-\frac{J^{2}}{G\Mbh a}\,,\nonumber \\
416: r_{p}\!=\! a(1-e)\,,\qquad &  & P\!=\!\frac{2\pi a^{3/2}}{\sqrt{G\Mbh}}\,.\label{eq:Kepler}\end{eqnarray}
417:  During most of the inspiral $1\!-\! e\!\ll\!1$, and the periapse
418: can be approximated by $r_{p}\!\approx\! J^{2}/2G\Mbh$. This remains
419: valid until the last phases of the inspiral. 
420: 
421: The prompt infall rate (Eq. \ref{e:Gp}) can be expressed in terms
422: of the maximal semi-major axis, \begin{equation}
423: \Gamma_{p}=\int_{0}^{a_{p}}\frac{\mathrm{d}aN_{\mathrm{iso}}(a)}{\ln(J_{m}/J_{lc})t_{r}(a)}\,.\label{eq:prompt}\end{equation}
424: 
425: 
426: We will assume Keplerian orbits throughout most of this paper except
427: for section (\ref{ss:MC}), where we employ the general relativistic
428: potential of the MBH.
429: 
430: 
431: \subsection{Slow inspiral}
432: 
433: \label{ss:inspiral}
434: 
435: The derivations of the conditions necessary for slow inspiral and
436: of the inspiral rate follow closely those of prompt infall, but with
437: two important differences. (1) The time to complete the inspiral is
438: not the infall time $P_{0}$, but rather $t_{0}\!\gg\! P_{0}$. (2)
439: There is no contribution from the full loss-cone regime, where stars
440: are scattered multiple times each orbit. This is because inspiral
441: in this regime would require that the \emph{very same} star that was
442: initially deflected into an eccentric orbit, be re-scattered back
443: into it multiple times. The probability for this happening is effectively
444: zero%
445: \footnote{This is to be contrasted with prompt infall from the full loss-cone
446: regime, where a star can reach the MBH by being scattered \emph{once}
447: into the loss-cone just before crossing $a_{p}$ toward the MBH. %
448: }.
449: 
450: In analogy to the radial scale $a_{p}$ of prompt infall, which delimits
451: the volume where stars can avoid scattering for a time $P_{0}\!<\! t_{J}$,
452: and thus maintain their infall orbit until they reach the MBH, the
453: inspiral criterion $t_{0}\!<\! t_{J}$ defines a critical radius $a_{c}\!\ll\! a_{p}$
454: (or equivalently, a critical energy $\varepsilon_{c}\!\gg\!\varepsilon_{p}$).
455: Stars starting the inspiral from orbits with $a_{0}\!<\! a_{c}$ ($\varepsilon_{0}\!>\!\varepsilon_{c}$)
456: will complete it with high probability, whereas stars starting with
457: $a_{0}\!>\! a_{c}$ ($\varepsilon_{0}\!<\!\varepsilon_{c}$), will
458: sample all $J$ values before they spiral in significantly, \emph{regardless
459: of $J_{0}$} and ultimately either (1) fall in the MBH, (2) diffuse
460: in energy to much wider orbits or (3) into the much tighter orbits
461: of the diffusive regime. Since we assume a steady state DF, outcomes
462: (2) and (3) represent a trivial, DF-preserving redistribution of stars
463: in phase space, which does not affect the statistical properties of
464: the system. Whether stars spiral in or fall in depends, statistically,
465: only on $a_{0}$ ($\varepsilon_{0}$). We use below Monte Carlo simulations
466: (\S\ref{ss:MC}, Fig. \ref{f:GWfrac}) to estimate the inspiral probability
467: function, $S(a_{0})$, which describes the probability of completing
468: inspiral when starting from an orbit with semi-major axis $a_{0}$
469: ($S\!\rightarrow\!1$ for $a_{0}\!\ll\! a_{c}$, $S\!\rightarrow\!0$
470: for $a_{0}\!\gg\! a_{c}$). The inspiral rate for stars of type $s$
471: with number fraction $f_{s}$ is then 
472: 
473: \begin{equation}
474: \Gamma_{i}=f_{s}\int_{0}^{\infty}\!\frac{\mathrm{d}aN(a)S(a)}{\ln(J_{m}/J_{lc})t_{r}(a)}\simeq f_{s}\int_{0}^{a_{c}}\!\frac{\mathrm{d}aN(a)}{\ln(J_{m}/J_{lc})t_{r}(a)}\,,\label{e:SU}\end{equation}
475: where roughly $S(a_{c})\!\sim\!0.5$. 
476: 
477: 
478: 
479: 
480: \section{Parameter dependence of the inspiral rate}
481: 
482: \label{s:inspscat}
483: 
484: 
485: 
486: In this section we derive some analytical results for the inspiral
487: rate. In order to keep the arguments transparent, we neglect relativistic
488: deviations from Keplerian motion. Relativistic orbits are discussed
489: in section (\ref{ss:MC}).
490: 
491: Consider a star of mass $\Ms$ orbiting a MBH of mass $\Mbh$ on a
492: bound Keplerian orbit with semi-major axis $a$ and angular momentum
493: $J$. When the star arrives at periapse, it loses some orbital energy
494: $\Delta E$ by GW emission. As a result, the orbit shrinks and its
495: energy increases. For highly eccentric orbits the periapse of the
496: star is approximately constant during inspiral in absence of scattering.
497: We define the inspiral time $t_{0}$ as the time it takes the initial
498: energy $\varepsilon_{0}$ to grow formally to infinity. If the energy
499: loss per orbit is constant, then for $e\rightarrow1$
500: 
501: \begin{equation}
502: t_{0}=\int_{\varepsilon_{0}}^{\infty}\frac{\mathrm{d}\varepsilon}{(\mathrm{d}\varepsilon/\mathrm{d}t)}\approx\frac{1}{\Delta E}\int_{\varepsilon_{0}}^{\infty}\mathrm{d}\varepsilon P(\varepsilon)=\frac{2\varepsilon_{0}P_{0}}{\Delta E},\end{equation}
503:  or \begin{equation}
504: t_{0}(r_{p},a)=\frac{2\pi\sqrt{GM_{\bullet}a}}{\Delta E}.\label{e:t0}\end{equation}
505: 
506: 
507: For GW, $\Delta E$ is given by (Peters \cite{Pe64}) 
508: 
509: \begin{equation}
510: \Delta E_{\mathrm{GW}}=\frac{8\pi}{5\sqrt{2}}f(e)\frac{\Ms c^{2}}{\Mbh}\left(\frac{r_{p}}{r_{S}}\right)^{-7/2}\,,\label{e:DEgw}\end{equation}
511: where\begin{equation}
512: f(e)=\frac{1+\frac{73}{24}e^{2}+\frac{37}{96}e^{4}}{(1+e)^{7/2}}\,,\label{e:f(e)}\end{equation}
513: and $r_S=2G\Mbh/c^2$ is the Schwarzschild radius. 
514: 
515: During all but the last stages of the inspiral $e\!\sim\!1$, in which
516: case $r_{p}/r_{S}=4(J/J_{lc})^{2}$ and
517: 
518: \begin{equation}
519: \Delta E_{\mathrm{GW}}=E_{1}(J/J_{lc})^{-7}\,,\qquad E_{1}\equiv\frac{85\pi}{3\!\times\!2^{13}}\frac{\Ms c^{2}}{\Mbh}\,.\label{e:DEgwJ}\end{equation}
520:  Gravity waves also carry angular momentum, \begin{equation}
521: \Delta J_{\mathrm{GW}}=-\frac{16\pi}{5}g(e)\frac{G\Ms}{c}\left(\frac{r_{p}}{r_{S}}\right)^{-2}\,,\label{e:DJgw}\end{equation}
522: \begin{equation}
523: g(e)=\frac{1+\frac{7}{8}e^{2}}{(1+e)^{2}}\,.\label{e:g(e)}\end{equation}
524: Generally, the change in $J$ in the course of inspiral is dominated
525: by two-body scattering, and $\Delta J_{\mathrm{GW}}$ can be neglected
526: until $a$ becomes very small. 
527: 
528: 
529: 
530: It is convenient to refer the timescales in the system to the relaxation
531: time at the MBH radius of influence,
532: 
533: \begin{equation}
534: t_{h}=A_{p}\left(\frac{\Mbh}{\Ms}\right)^{2}\frac{P(r_{h})}{N_{h}\mathrm{log}\Lambda_{1}}\,,\label{e:tr_rh}\end{equation}
535: where $\Lambda_{1}\!=\!\Mbh/\Ms$ $(r_{S}/r_{h})^{1/4}$ (Miralda-Escud\'{e}
536: \& Gould \cite{MG00}), and $A_{p}\!\simeq\!0.2$ for $p\!=\!0$
537: (Alexander \& Hopman \cite{AH03}). The relaxation time at any radius
538: is then
539: 
540: \begin{equation}
541: t_{r}(a)=t_{h}\left(\frac{a}{r_{h}}\right)^{p}\,,\label{e:tr}\end{equation}
542:  where we associate the typical relaxation time on an orbit with that
543:  at its semi-major axis. This is a good approximation, since
544:  theoretical arguments (Bahcall \& Wolf \cite{BW76}; \cite{BW77}) and
545:  simulations (Freitag \& Benz \cite{Fre02}; Baumgardt et
546:  al. \cite{Baum04a} \cite{Baum04b}; Preto et al. \cite{P04}) indicate
547:  that $0\lesssim p\lesssim0.25$, and so $t_{r}$ is roughly independent
548:  of radius. The angular momentum relaxation time is
549: 
550: \begin{equation}
551: t_{J}=\left[\frac{J}{J_{m}(a)}\right]^{2}\left(\frac{a}{r_{h}}\right)^{p}t_{h}\,.\label{e:tp}\end{equation}
552: 
553: 
554: 
555: 
556: Dissipational inspiral takes place in the presence of two-body scattering.
557: When $t_{0}\!\sim\! t_{J}$, both effects have to be taken into account.
558: It is useful to parametrize the relative importance of dissipation
559: and scattering by the dimensionless quantity
560: 
561: \begin{equation}
562: s(J,a)\equiv\frac{t_{0}(J,a)}{t_{J}(J,a)}=\left(\frac{a}{d_{c}}\right)^{3/2}\left(\frac{a}{r_{h}}\right)^{-p}\left(\frac{J}{J_{lc}}\right)^{5},\label{e:s}\end{equation}
563:  where we introduce the ($p$-independent) length scale\begin{equation}
564: d_{c}\equiv\left(\frac{8\sqrt{G\Mbh}E_{1}t_{h}}{\pi c^{2}}\right)^{2/3}\,,\label{e:D}\end{equation}
565:  which is of the same order as $a_{c}$ (Eq. \ref{e:ac}) and is $\ll\! r_{h}$.
566: We define some critical value $s_{\mathrm{crit}}\!\ll\!1$ such that
567: the inspiral is so rapid that the orbit is effectively decoupled from
568: the perturbations.
569: 
570: The three phases of inspiral can be classified by the value of $s$.
571: In the {}``scattering phase'' the star is far from the region in
572: phase space where GW emission is efficient and $s\!\gg\!1$. With
573: time it may scatter to a lower-$J$ orbit, enter the {}``transition
574: phase'', where $s\!\sim\!1$, and start to spiral in. If it is not
575: scattered into the MBH or to a wide orbit, it will eventually reach
576: the stage where $s\!<\! s_{\mathrm{crit}}$. It will then enter the
577: {}``dissipation phase'' where it spirals-in deterministically according
578: to equations (\ref{e:DEgw}--\ref{e:g(e)}). Note that eventually
579: the $e\!\rightarrow\!1$ approximation is no longer valid. 
580: 
581: Here we are mainly interested in understanding how the interplay between
582: two-body scattering and energy dissipation in the first two phases
583: sets the initial conditions for the GW emission in the final phase.
584: It should be emphasized that the onset of the dissipation phase does
585: not necessarily coincide with the emission of \emph{detectable} GW.
586: For example, while a star is well into the dissipation phase by the
587: time $s\!<\! s_{\mathrm{crit}}\!\sim\!10^{-3}$, the orbit has still
588: to decay substantially before the GW frequency becomes high enough
589: to be detected by \LISA. 
590: 
591: We derive an analytical order of magnitude estimate for the critical
592: semi-major axis $a_{c}$ by associating it with $s\!=\!1$ orbits
593: in the transition phase. Since $s$ falls steeply with $J$, we set
594: $J\!=\! J_{lc}$ and solve $ $$s(J_{lc},a_{c})\!=\!1$ for $a_{c}$,
595: obtaining 
596: 
597: \begin{equation}
598: \frac{a_{c}}{r_{h}}=\left(\frac{d_{c}}{r_{h}}\right)^{3/(3-2p)}.\label{e:ac}\end{equation}
599:  The MC simulations below (\S\ref{ss:MC}) confirm that this analytical
600: estimate corresponds within a factor of order unity to the semi-major
601: axis where the inspiral probability $S(a_{c})\!\sim\!0.5$, for a
602: wide range of masses (see table {[}\ref{t:stars}{]}). Expression
603: (\ref{e:SU}) for the rate can then be approximated by
604: 
605: \begin{equation}
606: \Gamma_{i}\sim\frac{f_{s}N_{h}}{t_{h}\ln[J_{m}(a_{c})/J_{lc}]}\left(\frac{a_{c}}{r_{h}}\right)^{3/2-2p},\label{e:Gammai}\end{equation}
607:  where $N_{h}$ is the number of stars within $r_{h}$. 
608: 
609: The dependence of the inspiral rate on $p$ (at fixed $d_{c}/r_{h}$
610: and neglecting the logarithmic terms) can be examined by writing $\Gamma_{i}\!\simeq\! g(p)f_{s}(N_{h}/t_{h})$,
611: where
612: 
613: \begin{equation}
614: g(p)=\left(\frac{d_{c}}{r_{h}}\right)^{\left.3(3-4p)\right/2(3-2p)}.\label{e:ggamma}\end{equation}
615: The pre-factor $g(p)$ grows with $p$ over the relevant range (see
616: also Ivanov \cite{IV02}). This reflects the fact that the inspiral
617: rate is determined by the number of stars within $a_{c}$, rather
618: than the total number of stars within $r_{h}$ . The concentration
619: of the cusp increases with $p$, so that there are more stars within
620: $a_{c}$. This result suggests qualitatively that in a mass-segregated
621: population, the heavier stars (higher $p$) will have an enhanced
622: GW event rate compared to the light stars (lower $p$).
623: 
624: >From equations (\ref{e:D}--\ref{e:ggamma}) it follows that \begin{equation}
625: \Gamma_{i}\propto t_{h}^{-2p/(3-2p)}.\end{equation}
626: Since $p\!\sim\!0$ for typical stellar cusps around MBHs, this means
627: that the inspiral rate is nearly independent of the relaxation time
628: for such cusps. This counter-intuitive result reflects the near balance
629: between two competing effects. When scattering is more efficient,
630: stars are supplied to inspiral orbits at a higher rate, but are also
631: scattered off them prematurely at a higher rate, so the volume of
632: the diffusive regime, which contributes stars to the inspiral ($\sim\! a_{c}^{3}$),
633: decreases. This is in contrast to the prompt disruption rate $\Gamma_{p}$,
634: which increases as the relaxation time becomes shorter%
635: \footnote{The prompt disruption rate is $\Gamma_{p}\!\sim\!\left.N(<\! a_{p})(r_{S}/a_{p})\right/P(a_{p})$
636: (Syer \& Ulmer \cite{SU99}, Eq, 10). For prompt disruption $a_{p}\!\sim\! t_{h}^{2/(5-2p)}$
637: (Alexander \& Hopman \cite{AH03}), and so the rate scales as $\Gamma_{p}\!\sim\! t_{h}^{-(2+2p)/(5-2p)}$
638: .%
639: }.  It then follows that enhanced scattering, such as by massive perturbers
640: (e.g. clusters, giant molecular clouds; Zhao, Haehnelt \& Rees \cite{ZHR02})
641: increases the prompt disruption rate, but will not enhance the rate
642: of inspiral events.
643: 
644: The dependence of the GW event rate on the mass of the MBH can be
645: estimated from Eq. (\ref{e:Gammai}) and the empirical $M_{\bullet}$--$\sigma$
646: relation \begin{equation}
647: M_{\bullet}=1.3\times10^{8}M_{\odot}\left(\frac{\sigma}{200\,\mathrm{km\, s^{-1}}}\right)^{\beta},\label{e:Msigma}\end{equation}
648:  where $\beta\!\sim\!4$ (Ferrarese \& Merritt \cite{FM00}; Gebhardt
649: et al. \cite{Geb00}; Tremaine et al. \cite{Tr02}). Note that the
650: $\Mbh$--$\sigma$ relation implies that the stellar number density
651: at the radius of influence, $n_{h}$ is larger for lighter MBHs: for
652: example, for $\beta\!=\!4$, $n_{h}\!\propto\! N_{h}r_{h}^{-3}\!\propto\!\Mbh^{-1/2}$,
653: where we assumed that $N_{h}\!\propto\!\Mbh$; the consequences of
654: this for the dependence of the rate for prompt tidal disruptions on
655: the MBH mass were discussed by Wang \& Merritt (\cite{WM04}).
656: 
657: The GW event rate depends on $\Mbh$ as 
658: 
659: \begin{equation}\label{eq:massdependence}
660: \Gamma_{i}\propto\Mbh^{3/\beta-1}\,.\label{e:massdependence}\end{equation}
661: This dependence is weak, e.g. $\Gamma_{i}\!\propto\!\Mbh^{-1/4}$for
662: $\beta=4$. Thus, the rate becomes \textit{higher} for lower mass
663: MBHs. If $\Mbh\sim10^{3}\,\Mo$ IBHs indeed exist in stellar clusters
664: and the $M_{\bullet}$--$\sigma$ relation can be extrapolated to
665: these masses, then they may be more likely to capture stars than MBHs.
666: This, however, does not necessarily translate into more GW sources.
667: The strain $h$ of GW decreases with the mass of the MBH, so that
668: these sources have to be closer by in order to observe their GW emission.
669: Another restriction is that for IBHs the tidal force is so strong
670: that white dwarfs are tidally disrupted well outside the event horizon,
671: which precludes them from being \emph{\LISA} sources. These issues
672: are further discussed in \S\ref{s:results}.
673: 
674: 
675: \section{Orbital evolution with dissipation and scattering}
676: 
677: \label{s:approaches} We present three different methods for analyzing
678: inspiral in the presence of scattering. The first approach is based
679: on Monte Carlo (MC) simulations, which follow a star on a relativistic
680: orbit, described by $\varepsilon$ and $J$, and add small perturbations
681: to simulate energy dissipation and random two-body scattering. The
682: second approach consists of direct numerical integration of the time
683: dependent diffusion-dissipation equation. The third approach is a
684: heuristic semi-analytical effective model that can describe the {}``effective''
685: trajectory of a star through phase space, as well as the statistical
686: properties of an ensemble of such trajectories.
687: 
688: 
689: 
690: The three approaches are complementary. The MC simulations allow a
691: direct realization of the micro-physics of the system, since they
692: follow the perturbed orbits of individual stars. They also offer much
693: flexibility in setting the initial conditions of the numerical experiments,
694: but the results are subject to statistical noise and are not easy
695: to generalize. The diffusion-dissipation equation on the other hand
696: deals directly with the DF, and allows an analytical formulation of
697: the problem in terms of partial differential equations, which are
698: solved numerically. However, computational limitations do not allow
699: covering as large a dynamical range as in the MC simulations. Finally,
700: the heuristic effective model has the advantage of its intuitive directness
701: and relative simplicity of use. We find that all three methods give
702: the same results for the same underlying assumptions (Fig. \ref{f:Jlossdist}).
703: This inspires confidence in the robustness of the analysis.
704: 
705: To compare the three methods, we stop the simulation at the point
706: where $s=s_{{\rm crit}}=10^{-3}$, and we plot the DF of the angular
707: momenta of the stars. From that point on the stars are effectively
708: decoupled from the cluster and can spiral in undisturbed. We then
709: use the MC method, which can be easily extended to follow the stars
710: in the dissipation phase (section \ref{s:results}), to find the DF
711: of the eccentricities of stars which enter the \emph{\LISA} band.
712: 
713: 
714: 
715: 
716: \subsection{Monte Carlo simulations\label{sub:MC}}
717: 
718: \label{ss:MC}
719: 
720: The MC simulations generally follow the scheme used by Hils \& Bender
721: (\cite{HB95}) to study the event rate of GW. The star starts on an
722: initial orbit with $(\varepsilon_{0},J_{0})$ such that dissipation by
723: GW emission is negligible. The initial value of $J$ is not of
724: importance as the angular momentum is quickly randomized in the first
725: few steps of the simulation. Every orbital period $P(\varepsilon)$,
726: the energy and angular momentum are modified by $\delta\varepsilon$
727: and $\delta J$, and the orbital period and periapse are recalculated
728: (this diffusion approach is justified as long as $P\!\ll\! t_{J}$).
729: The simulation stops when the star decays to an orbit with $s\!<\!
730: s_{\mathrm{crit}}\!=\!10^{-3}$ or when $J\!<\! J_{lc}$ and it falls in
731: the MBH (escape to a less bound orbit is not an option here since
732: energy relaxation is neglected and only dissipation is considered; see
733: \S\ref{ss:inspiral}).
734: 
735: When stars reach high eccentricities as a result of scattering, the
736: periapse approaches the Schwarzschild radius to the point where the
737: Newtonian approximation breaks down.The MC simulations take this into
738: account by integrating the orbits in the relativistic potential of
739: a Schwarzschild MBH. The periapse of the star moving in a Schwarzschild
740: spacetime is related to its angular momentum by one of the three roots
741: of the equation\begin{equation}
742: E_{{\rm GR}}^{2}=\left(c^{2}-\frac{2GM_{\bullet}}{r}\right)\left(c^{2}+\frac{J^{2}}{r^{2}}\right)\equiv V_{{\rm GR}}(r,J).\label{eq:GRpot}\end{equation}
743: The term on left hand side of this equation is the squared specific
744: relativistic energy of the star (including its rest mass), while the
745: right hand side is the effective GR potential. A star on a bound orbit
746: is not captured by the MBH as long as equation (\ref{eq:GRpot}) has
747: three real roots for $r$. The smallest root is irrelevant for our
748: purposes. The intermediate root (the turning-point) is the periapse
749: $r_{p}$ of the orbit, and the largest root is the apo-apse $r_{a}$.
750: The semi-major axis $a$ and eccentricity $e$ are defined by $a=(r_{p}+r_{a})/2$
751: and $e=1-r_{p}/a$ (see Eq. \ref{eq:Kepler}). For bound orbits the
752: loss cone $J_{lc}$ is a very weak function of energy and is well
753: approximated by Eq. (\ref{e:Jlc}).
754: 
755: For given values of $E$ and $J$, the GR periapse is smaller than
756: the Newtonian one, and therefore the orbits can be more eccentric,
757: and the dissipation can be stronger than implied by the Newtonian
758: approximation. The Keplerian relations between energy, angular momentum
759: and the orbital parameters (Eqs \ref{eq:Kepler}) are replaced by 
760: 
761: \begin{equation}
762: E_{{\rm GR}}^{2}=\frac{(q-2-2e)(q-2+2e)}{q(q-3-e^{2})};\label{eq:Egr}\end{equation}
763: \begin{equation}
764: J^{2}=\frac{q^{2}}{q-3-e^{2}}\left(\frac{GM_{\bullet}}{c}\right)^{2},\label{eq:Jgr}\end{equation}
765: where $q=2(1-e^{2})a/r_{S}$ (Cutler, Kennefick \& Poisson
766: \cite{CKP94}).  The condition for a non-plunging orbit can also be
767: expressed in terms of the eccentricity and the semi-major axis,
768: $2(1-e^{2})a=(6+2e)r_{S}$ (Cutler et al. \cite{CKP94}). This
769: corresponds to a maximal eccentricity for a star on a non-plunging
770: orbit\begin{equation} e_{{\rm
771: max}}(a)=-\frac{r_{S}}{2a}+\sqrt{(r_{S}/2a)^{2}-3r_{S}/a+1},\label{e:eL}\end{equation}
772: which increases with $a$. If
773: $a_{L}=(P_{L}^{2}/4\pi^{2}GM_{\bullet})^{1/3}$ is the maximal
774: semi-major axis a star may have in order to be detectable by \LISA\ ,
775: the maximal eccentricity of a star detectable by \LISA\ is
776: $e_{\max}(a_{L})$.
777: 
778: The step in $J$-space per orbit is the sum of three terms \begin{equation}
779: \delta J(\varepsilon,J)=\Delta_{1}J_{{\rm scat}}(\varepsilon,J)+\chi\Delta_{2}J_{\mathrm{scat}}(\varepsilon,J)-\Delta J_{\mathrm{GW}}(\varepsilon,J)\,.\label{e:dJ}\end{equation}
780:  The first and second terms represent two-body scattering (Eqs. \ref{e:tJ},\ref{eq:Jcoeffs},\ref{e:diffcoeff}),
781: with%
782: \footnote{The drift term $\Delta_{1}J_{{\rm scat}}$ represents a bias to scatter
783: \emph{away} from the MBH due to the 2D character of the direction
784: of the velocity vector $\mathbf{v}$. This can be expressed geometrically
785: by considering a circle of radius $\Delta v_{\mathrm{scat}}$ centered
786: on $\mathbf{v}$ ($\mathbf{v}\!+\!\Delta\mathbf{v}{}_{{\rm scat}}$is
787: the change per orbit due to scattering). This circle is intersected
788: by a second circle of radius $v$ that passes through $\mathbf{v}$
789: and is centered on the radius vector to the MBH. The section of the
790: first circle that leads away from the MBH is slightly larger than
791: the section leading toward it. %
792: } $\Delta_{1}J_{{\rm scat}}\!=\!\left\langle \Delta J\right\rangle P\!=\! J_{m}^{2}P(\varepsilon)/(2t_{r}J)$
793: and $\Delta_{2}J_{\mathrm{scat}}\!=\!\sqrt{\left\langle \Delta J^{2}\right\rangle P}\!=\!\sqrt{P(\varepsilon)/t_{r}}J_{m}(\varepsilon)\!=\!\sqrt{P(\varepsilon)/t_{J}}J$.
794: The random variable $\chi$ takes the values $\pm1$ with equal probabilities.
795: The third term is the deterministic angular momentum loss due to GW
796: emission (Eq. \ref{e:DJgw}).The energy step per orbit is deterministic
797: (diffusion in energy is neglected)
798: 
799: \begin{equation}
800: \delta E(\varepsilon,J)=\Delta
801: E_{\mathrm{GW}}(\varepsilon,J)\,.\label{e:dE}\end{equation} In
802: order to increase the speed of the simulation we use an adaptive
803: time-step. We checked for some cases that taking a smaller time-step
804: does not affect the results.
805: 
806: The DF of inspiraling stars is generated by running many simulations
807: (typically $3\!\times\!10^{4}$) \texttt{\textbf{}}of stellar
808: trajectories through phase space with the same initial value for $J_0$
809: for given initial semi-major axis $a_0$, but with different random
810: perturbations. We verified that the initial value of $J_0$ is
811: irrelevant, as long as $J_0\gg J_{lc}$. We record the fraction of
812: stars that avoid falling in the MBH, $S(a_{0})$, and the value of $J$
813: at $s_{\mathrm{crit}}$ for those stars that reach the
814: dissipation-dominated phase, thereby obtaining the DF
815: $W(J;a_{0})$. This is repeated for a range of $a_{0}$ values. The
816: integrated DF over all cusp stars, $W(J)$, is then obtained by taking
817: the average of all DFs, weighted by $N(a_{0})S(a_{0})\mathrm{d}a_{0}$
818: (cf Eq. \ref{e:SU}).
819: 
820: 
821: \subsection{Diffusion equation with GW dissipation}
822: 
823: \label{ss:DiffDiss} 
824: 
825: The DF of the inspiraling stars at the onset of the dissipation dominated
826: phase can be obtained directly by solving the diffusion-dissipation
827: equation. This approach was taken by Ivanov (\cite{IV02}), who included
828: a GW dissipation term and obtained analytic expressions for the GW
829: event rate for $J\!\gg\! J_{lc}$ under various simplifying assumptions.
830: Here we are interested DF of stars very near the loss-cone, and so
831: we integrate the diffusion-dissipation equation numerically with two
832: simplifying assumptions. (1) We neglect the drift term in the diffusion-dissipation
833: equation. This can be justified by noting that the drift grows linearly
834: with time, $\delta J_{1}(t)=J_{m}^{2}t/(2t_{r}J)$, while the change
835: due to the random walk grows as $\delta J_{2}(t)=(t/t_{r})^{1/2}J_{m}$.
836: For a star with initial angular momentum $J$ the drift becomes important
837: only for $t\!>\! t_{{\rm drift}}=4(J/J_{m})^{2}t_{r}\!=\!4t_{J}$.
838: Since we are interested in the timescales $t\!\sim\! t_{0}\le\! t_{J}$,
839: the drift is only a small correction. (2) We assume Keplerian orbits.
840: These approximations are validated by the very good agreement with
841: the MC simulation results (section \ref{ss:compare}).
842: 
843: With these two assumptions, the diffusion-dissipation is (cf Eqs.
844: \ref{e:FP}, \ref{e:Fss})
845: 
846: \begin{eqnarray}
847: \frac{\partial N(\varepsilon,J;t)}{\partial t} & = & \frac{J_{m}^{2}(\varepsilon)}{2t_{r}}\frac{\partial^{2}}{\partial J^{2}}N(\varepsilon,J;t)\nonumber \\
848:  & + & \frac{\partial}{\partial\varepsilon}\left[\dot{E}_{\mathrm{GW}}(J)N(\varepsilon,J;t)\right],\label{e:diffdiss}\end{eqnarray}
849:  where $\dot{E}_{\mathrm{GW}}=\Delta E_{\mathrm{GW}}/P$ is the rate
850: at which energy is lost to GW. The first term accounts for diffusion
851: in $J$-space and the second represents the energy dissipated by GW
852: emission. As with the MC simulations, the diffusion coefficient $\left\langle \Delta J^{2}\right\rangle \!=\! J_{m}^{2}/t_{r}$
853: is an input parameter rather than resulting from a self-consistent
854: calculation.
855: 
856: 
857: 
858: The initial conditions consist of an isotropic cusp DF, $N(\varepsilon,J)\propto\varepsilon^{p-5/2}J$
859: and the boundary condition that the DF vanish on the loci $s\!=\!\scr\!=\!10^{-3}$
860: and $J\!=\! J_{lc}$ in the ($\varepsilon$,$J$)-grid. The initial
861: DF is evolved in time over the $(\varepsilon,J)$-grid until $t\!\sim\! t_{J}$,
862: when a relaxed steady state is achieved. The integration is done using
863: a Forward Time Centered Space (FTCS) representation of the diffusion
864: term, and an 'upwind' differential scheme for the dissipation term
865: (see e.g. Press et al. \cite{PTVF77}). After each time step (which
866: are chosen small enough to obey the Courant condition) the DF is re-normalized
867: so that captured stars are replenished.
868: 
869: After relaxation, the DF is non-zero up to the boundary $s\!=\!\scr$,
870: as stars are redistributed in phase space by diffusion at large $J$
871: and by dissipation at small $J$. The DF at $s_{\mathrm{crit}}$ is
872: then extracted to construct the angular momentum DF, $W(J)$. Although
873: the stars continue their trajectory in phase space beyond $s_{\mathrm{crit}}$
874: all the way to the last stable orbit, this does not change $W(J)$
875: because the rapid energy dissipation does not allow much $J$-diffusion.
876: 
877: 
878: \subsection{Analytical model of effective orbit }
879: 
880: \label{s:model}
881: 
882: Stars follow complicated stochastic tracks in $(J,a)$ space; due
883: to scattering they move back and forth in $J$, while they always
884: drift to smaller $a$ due to GW dissipation (energy diffusion by scattering
885: is neglected). The drift rate depends strongly on $J$. Figure (\ref{f:Ja})
886: shows some examples of stellar tracks in phase space, taken from the
887: MC simulations.
888: 
889: %
890: \begin{figure}
891: %%%%\includegraphics[%
892: %%%%  scale=0.5]{f1.EPS}
893: 
894: \plotone{f1.eps}
895: 
896: \figcaption[FileName]{\label{f:Ja}Examples of tracks of stars from the MC simulation (solid
897: lines), and a track from the effective model (dashed line) with $\xi=1$.
898: The effective track represents the stochastic tracks well. The tracks
899: extend to the point where $s\!=\!10^{-3}$. The relevant parameters
900: are given in the first line of table (\ref{t:stars}), i.e. inspiral
901: of a white dwarf into a MBH.}
902: \end{figure}
903: 
904: 
905: In this section we propose a heuristic analytical model, which captures
906: some essential results of the MC simulations, while providing a more
907: intuitive understanding.
908: 
909: In this model we define the equations of motion of an {}``effective
910: track'' of a star, which is deterministic and can be solved analytically.
911: With this method we follow stars that start at some given initial
912: semi-major axis $a_{0}$ and angular momentum $J_{0}$, and follow
913: the star during its first two stages of inspiral, i.e., until the
914: moment that $s\!<\! s_{\mathrm{crit}}$.
915: 
916: The DF of stars which reach $s\!=\! s_{\mathrm{crit}}$ is determined
917: by the evolution of the orbital parameters ($J,a$) in the region
918: where energy dissipation is efficient. Because of the strong dependence
919: of the energy $\Delta E$ which is dissipated per orbit on angular
920: momentum (or, equivalently, periapse), this is a small region $\Delta J$
921: in $J$-space, the size of which is of order of the loss-cone, $\Delta J\!\sim\! J_{lc}$.
922: The time $\Delta t$ it typically takes a star with semi-major axis
923: $a_{0}$ performing a random walk in $J$ to cross this region is
924: $\Delta t\!=\![\Delta J/J_{m}(a_{0})]^{2}t_{h}$ ($t_{r}\!=\! t_{h}$
925: assumed). This can be used to introduce an effective $J$-{}``velocity''
926: $\Delta J/\Delta t\!=\!-J_{m}^{2}(a_{0})/J_{lc}t_{h}$.
927: 
928: The effective velocity of the semi-major axis is given by
929: 
930: \begin{equation}
931: \frac{\mathrm{d}a(J,a)}{\mathrm{d}t}=\frac{da}{dE}\dot{E}=-\frac{2a^{2}}{G\Mbh}\frac{\Delta E_{\mathrm{GW}}(J)}{P(a)}.\label{e:aeff}\end{equation}
932: 
933: 
934: These two equation define a deterministic time evolution in ($J,a$)
935: space from given initial values ($J(0)\!=\! J_{0},\, a(0)\!=\! a_{0}$),
936: to a final point ($J_{f},a_{f}$), where $s\!=\!\scr$. To recover
937: $W(J;a_{0})$ at $s_{\mathrm{crit}}$, we introduce some scatter in
938: the effective velocities $\mathrm{d}J/\mathrm{d}t$,
939: 
940: \begin{equation}
941: \frac{\mathrm{d}j}{\mathrm{d}t}=-\xi y\frac{j_{m}^{2}(a_{0})}{t_{h}},\label{e:Jeff}\end{equation}
942:  where $j$ denotes angular momentum in terms of $J_{lc}$, $y$ is
943: drawn from the positive wing of the normalized Gaussian distribution,
944: $G_{1}(y)$, and $\xi\!\sim\!1$ (a free parameter) is the width of
945: the distribution, which can vary depending on the system's parameters.
946: Equations (\ref{e:aeff},\ref{e:Jeff}) can be solved analytically,
947: 
948: \begin{equation}
949: j(t)\!=\! j_{0}\!-\! y\xi j_{m}^{2}(a_{0})\frac{t}{t_{h}}\,,\qquad a(t)\!=\!\left[1\!-\!\frac{(d_{c}/a_{0})^{3/2}}{6\xi yj^{6}}\right]^{2}a_{0}\,.\label{e:at}\end{equation}
950:  The stopping condition $s(j,a)=s_{\textrm{crit}}$ gives an additional
951: relation between $j$ and $a$, so that the final values are related
952: by $a_{f}=\scr^{2/3}j_{f}^{-10/3}d_{c}.$ This ties the initial and
953: final values through the equations of motions
954: 
955: 
956: 
957: \begin{equation}
958: y(j_{f})=\frac{(6\xi)^{-1}(d_{c}/a_{0})^{3/2}j_{f}^{-6}}{1-\scr^{1/3}(d_{c}/a_{0})^{1/2}j_{f}^{-5/3}}.\label{e:bJ}\end{equation}
959: 
960: 
961: The relation between the initial distribution of $j$-velocities and
962: the final $j$ distribution at is $W(j_{f};a_{0})=G_{1}(y)[\mathrm{d}y(j_{f})/\mathrm{d}j_{f}]\,.$
963: The integrated DF over all the cusp stars is
964: 
965: \begin{equation}
966: W(j_{f})\mathrm{d}j_{f}=\frac{\int^{a_{c}}N(a)W(j_{f};a)\,\mathrm{d}a}{\int^{a_{c}}N(a)\,\mathrm{d}a}\,.\label{e:WJ}\end{equation}
967: 
968: 
969: 
970: \subsection{Comparison of the different methods}
971: 
972: \label{ss:compare}
973: 
974: We compare the results of the MC simulation, the diffusion-dissipation
975: equation and an integration of equations (\ref{e:Jeff}) and (\ref{e:aeff}).
976: In all cases the calculation is stopped when $s\!=\!10^{-3}$; at
977: that point the eccentricity is still approximately unity, but scattering
978: becomes entirely negligible so that even in the MC simulation the
979: quantities evolve essentially deterministically according to equations
980: (\ref{e:DEgw}-\ref{e:g(e)}). The calculations assume a cusp with
981: slope $p\!=\!0$, $\Mbh\!=3\!\times\!10^{6}\,\Mo$ and $\Ms\!=\!0.6$
982: (corresponding to WDs).
983: 
984: %
985: \begin{figure}
986: %\includegraphics[%
987: %  width=1.0\columnwidth,
988: %  keepaspectratio]{f2.EPS}
989: \plotone{f2.eps}
990: 
991: \figcaption[FileName]{\label{f:Jlossdist}A comparison of the DFs of angular momenta for
992: inspiraling stars in a $p\!=\!0$ stellar cusp, derived from (1) MC
993: simulations, (2) direct integration of the diffusion-dissipation equation
994: (Eq. \ref{e:diffdiss}) and (3) the effective model (Eq. \ref{e:WJ},
995: with $\xi\!=\!0.6$). The DFs are normalized over the displayed $J$-range.
996: The DFs are shown at $s\!=\!10^{-3}$, when the dynamics are dominated
997: by dissipation and scattering no longer plays a role, but the orbits
998: are still well outside the \LISA\ band, $P\!\gg\! P_{L}$ and $e\!\rightarrow\!1$.
999: This DF sets the initial conditions for the dissipation phase. The
1000: DF of the isotropic parent population (no sink, no dissipation; Eq.
1001: \ref{e:Niso}) is also shown for contrast. Unlike the DF of the inspiraling
1002: stars, it is dominated by high-$J$ stars.}
1003: \end{figure}
1004: 
1005: 
1006: Figure (\ref{f:Jlossdist}) shows the very good correspondence between
1007: the three approaches, whose underlying assumptions and approximations
1008: are summarized in table (\ref{t:comp}). One important conclusion
1009: is that all stars enter the dissipation phase with very small angular
1010: momenta, $1\!\lesssim\! J/J_{lc}\!\lesssim2$. For a suitable choice
1011: of $\xi\!\sim\!1$, the actual complicated stochastic tracks are well
1012: mimicked by the tracks of the {}``effective stars''. The semi-analytical
1013: approach not only identifies the correct scale of angular momentum
1014: of inspiraling stars, but reproduces the DF. Its practical value lies
1015: in the relative ease of its application compared to the time consuming
1016: MC simulations or the integration of the diffusion-dissipation equation. 
1017: 
1018: %\clearpage
1019: 
1020: \begin{table}
1021: 
1022: \caption{\label{t:comp}Comparison of assumptions in the 3 methods}
1023: 
1024: \begin{center}\begin{tabular}{lcccc}
1025: \hline 
1026: Method&
1027: GR potential&
1028:  $\Delta_{1}J_{{\rm scat}}$&
1029:  $\Delta J_{{\rm GW}}$&
1030:  Stop at $s_{{\rm crit}}$\tabularnewline
1031: \hline 
1032:  Monte Carlo&
1033: yes&
1034:  yes &
1035:  yes &
1036:  no$^a$\tabularnewline
1037:  Diffusion/Dissipation&
1038: no&
1039: no&
1040:  no &
1041:  yes\tabularnewline
1042:  Effective track&
1043: no&
1044: no&
1045:  no&
1046:  yes\tabularnewline
1047: \hline
1048: \multicolumn{5}{l}{$^{a}$ For the purpose of comparison to the other two methods,}\tabularnewline
1049: \multicolumn{5}{l}{ the MC simulation was terminated at $s_{\rm crit}$.}
1050: \tabularnewline
1051: \hline
1052: \end{tabular}\end{center}
1053: \end{table}
1054: %\centerline{\rule{0.99\columnwidth}{2pt}}
1055: 
1056: %\clearpage
1057: 
1058: \section{Inspiral rates and distribution of orbital parameters}
1059: 
1060: \label{s:results} 
1061: 
1062: We now proceed to apply the MC simulation to different types of COs
1063: inspiraling into a $3\!\times\!10^{6}\,\Mo$ MBH in a galactic center,
1064: and into a $10^{3}\,\Mo$ IBH in a stellar cluster. We find from the
1065: simulations the critical semi-major axis $a_{c}$ and calculate the
1066: DF of the eccentricities of \LISA\   sources. We stop the simulations
1067: when the orbital period falls below the longest period detectable
1068: by \LISA\  , $P_{L}\!=\!10^{4}\,\mathrm{s}$, which corresponds to
1069: a semi-major axis of 
1070: 
1071: \begin{equation}
1072: a_{L}=23.5M_{6}^{-2/3}r_{S},\label{e:aL}\end{equation}
1073: where $\Mbh=10^{6}M_{6}\,\Mo$.We record the eccentricity at that
1074: point and construct the DF of the eccentricity at $P_{L}$. It is
1075: straightforward to integrate the orbits in the eccentricity histograms
1076: backward (forward) in time to larger (smaller) values of $P$, see
1077: e.g. Barack \& Cutler (\cite{BC04a}).
1078: 
1079: The stars are distributed according to a powerlaw distribution
1080: with different values for $p$. The total number of stars within $r_h$
1081: was assumed to be $N_h=2\Mbh/\Mo$, with different number fractions for
1082: the respective species. See table (\ref{t:stars}) for the assumed
1083: parameters of the stellar populations.
1084: 
1085: 
1086: \subsection{Massive black holes in galactic centers}
1087: 
1088: \label{s:MBH}
1089: 
1090: We assume a $\Mbh\!=\!3\!\times\!10^{6}\,\Mo$ MBH as a representative
1091: example and make the simplifying assumption that the MBH is not
1092: spinning.  Real MBHs probably have non-zero angular momentum (e.g. see
1093: for evidence of spin of the MBH in our Galactic Center Genzel et
1094: al. \cite{GS03}). An important qualitative difference is that in the
1095: Schwarzschild case the eccentricity of a GW-emitting star decreases
1096: monotonically with time, which this is not so for non-zero angular
1097: momentum (e.g. Glampedakis et al. \cite{GHK02}).  We conducted several
1098: MC simulations with Kerr metric orbits and verified that in spite of
1099: changes in details, our overall results hold.
1100: 
1101: The tidal field of MBHs in this mass range disrupts main sequence
1102: stars before they can become detectable \LISA\ sources. A possible
1103: exception is our own Galactic Center, where the weak GW emission from
1104: very low mass main sequence stars (which are the densest and so the
1105: most robust against tidal disruption) may be detected because of their
1106: proximity (Freitag \cite{FR03}). However, here we only consider GW
1107: from COs. Table (\ref{t:stars}) summarizes the parameters assumed
1108: for the properties of white dwarfs (WDs), neutron stars (NSs) and
1109: stellar mass black holes (BHs). The BHs are assumed to be strongly
1110: segregated in a steep cusp because of their much heavier mass (Bahcall
1111: \& Wolf \cite{BW77}). The total (dark) mass in COs near MBHs is not
1112: known, but in future it may be constrained by deviations from Keplerian
1113: motions of luminous stars very near the MBH in the Galactic center
1114: (Mouawad et al. \cite{MO04}).
1115: 
1116: %\clearpage
1117: 
1118: \begin{table}
1119: 
1120: \caption{\label{t:stars} Parameters for stellar populations and inspiral rates}
1121: 
1122: \begin{center}\begin{tabular}{ccccccccc}
1123: \hline 
1124: Star &
1125: $\Ms$&
1126:  $t_{h}$&
1127: $r_{h}$&
1128:  $p$&
1129:  $f_{s}$&
1130: $a_{c}^{c}$&
1131: $a_{c}^{d}$&
1132:  $\Gamma_{i}$\tabularnewline
1133: &
1134:  {[}$\Mo${]}&
1135:  {[}Gyr{]}&
1136: {[}pc{]}&
1137: &
1138: &
1139: {[}mpc{]}&
1140: {[}mpc{]}&
1141:  {[}$\mathrm{Gyr^{-1}}${]}\tabularnewline
1142: \hline
1143: WD$^{a}$&
1144:  0.6 &
1145:  $10$ &
1146: 2&
1147:  0 &
1148:  $0.1$&
1149: 30&
1150: 20&
1151:  7.8 \tabularnewline
1152: NS$^{a}$&
1153:  1.4 &
1154: $5$&
1155: 2&
1156:  0 &
1157:  0.01&
1158: 40&
1159: 20&
1160:  1.8 \tabularnewline
1161: BH$^{a}$&
1162:  10 &
1163: $1$&
1164: 2&
1165:  1/4 &
1166:  $1(-3)$&
1167: 20&
1168: 10&
1169:  4.7 \tabularnewline
1170: NS$^{b}$&
1171:  1.4 &
1172: $1(-3)$ &
1173: 0.05&
1174:  0 &
1175:  0.01&
1176: 2&
1177: 1&
1178:  4.3 \tabularnewline
1179: BH$^{b}$&
1180:  10 &
1181: $1(-3)$ &
1182: 0.05&
1183:  1/4 &
1184:  1(-3)&
1185: 5&
1186: 2&
1187:  6.7 \tabularnewline
1188: \hline
1189: \multicolumn{9}{l}{$^{a}$ MBH with $\Mbh\!=\!3\times10^{6}\,\Mo$}\tabularnewline
1190: \multicolumn{9}{l}{$^{b}$ IBH with $\Mbh\!=\!10^{3}\,\Mo$.  }\tabularnewline
1191: %\hline
1192: \multicolumn{9}{l}{$^{c}$From equation (\ref{e:ac})}\tabularnewline
1193: %\hline
1194: \multicolumn{9}{l}{$^{d}$From the MC simulations}\tabularnewline
1195: \hline
1196: \end{tabular}\end{center}
1197: \end{table}
1198: 
1199: %\clearpage
1200: 
1201: Fig. (\ref{f:GWfrac}) shows the normalized inspiral probability
1202: function $S_{s}(a_{0})$ , where $s$ stands for WD, NS, and BH. The
1203: critical semi-major axis is similar for WDs and NSs, but much smaller
1204: for BHs, because of the smaller relaxation time and the higher central
1205: concentration due to mass segregation. The functions $S_{s}(a)$ are
1206: used to calculate the inspiral rates (Eq. \ref{e:SU}) that are listed
1207: in table (\ref{t:stars}).  The rate for WDs is highest, $\Gamma_{\rm
1208: WD}=7.8\,{\rm Gyr^{-1}}$. We find that because of mass segregation,
1209: the rate for stellar BHs is of the same order of magnitude as that for
1210: WDs, $\Gamma_{\rm BH}=4.7\,{\rm Gyr^{-1}}$ although their number
1211: fraction is lower by two orders of magnitude.  The hierarchy of rates
1212: in table (\ref{t:stars}) is not, of course, necessarily the same as
1213: cosmic rates that \LISA\ will observe; NSs and BHs are more massive
1214: than WDs, and can be observed at larger distances.
1215: 
1216: %
1217: \begin{figure}
1218: %\includegraphics[%
1219: %  width=1.0\columnwidth,
1220: %  keepaspectratio]{f3.EPS}
1221: 
1222: \plotone{f3.eps}
1223: 
1224: \figcaption[FileName]{\label{f:GWfrac}Dependence of the normalized fraction of inspiral
1225: events on initial semi-major axis, for the case of a MBH of $\Mbh\!=\!3\times10^{6}\,\Mo$ (asterisks)
1226: and an IBH (triangles). The solid line is for white dwarfs, the dashed
1227: lines for neutron stars, and the dotted lines for stellar mass black
1228: holes. For the parameters of these cases see table (\ref{t:stars}).}
1229: \end{figure}
1230: 
1231: 
1232: Fig. (\ref{f:WeMBH}) shows the eccentricity DFs of stars on orbits
1233: with $P\!=\! P_{L}$; note that since $a\!=\! a_{L}$ is fixed, the
1234: orbits are fully determined by $e$. The DFs show a strong bias to
1235: large eccentricities. The maximal eccentricity possible for a star
1236: orbiting a $3\!\times\!10^{6}\,\Mo$ MBH in the \emph{\LISA} band
1237: is $e_{\textrm{max}}\!=\!0.81$ (Eq. \ref{e:eL}). 
1238: 
1239: It should be emphasized that the histogram in figure (\ref{f:WeMBH})
1240: can be obtained deterministically from the DF $W(J;s_{\mathrm{crit}}\!=\!10^{-3})$
1241: in figure (\ref{f:Jlossdist}), because the stars have already reached
1242: the point where scattering is negligible. This would not be the case
1243: had the MC simulation been terminated at $s_{\mathrm{crit}}=1$ (e.g.
1244: Freitag \cite{FR03}), since then significant scatterings would continue
1245: to redistribute the orbital parameters. We find however that the final
1246: DF (at $P\!=\! P_{L}$), which is obtained by integrating $W(J;s_{\mathrm{crit}}\!=\!1)$
1247: forward in time according to the GW dissipation equations (Eqs. \ref{e:DEgw},\ref{e:DJgw})
1248: without scattering, is not much different from that shown in Fig.
1249: (\ref{f:WeMBH}). This coincidence is due to the fact that the stars
1250: with the largest eccentricities, which typically drop into the MBH,
1251: are replenished by the stars with slightly lower eccentricities. The
1252: main consequence of choosing $\scr$ too large is an overestimate
1253: of the total event rate; stars which actually fall into the MBH are
1254: erroneously counted as contributing to the GW event rate. Incidentally,
1255: even though stars that do not complete the inspiral are not individually
1256: resolvable, they will contribute to the background noise in the \LISA\ band
1257: (Barack \& Cutler \cite{BC04b}). 
1258: 
1259: A premature neglect of the effects of scattering in previous studies
1260: (e.g. Freitag \cite{FR01}) probably explains in part why our derived
1261: rates are significantly lower. Those studies usually assumed that
1262: stars will spiral in without further perturbations once $s\!=\!1$. We
1263: ran a simulation that was stopped at $s_{\mathrm{crit}}\!=\!1$ instead
1264: of $s_{\mathrm{crit}}\!=\!10^{-3}$. The event rate in that unrealistic
1265: case is about $\sim\!6$ times higher. This does not explain all of the
1266: discrepancies, which are hard to track as different methods are
1267: used. The different stopping criterion may also explain why our rates
1268: are lower than those estimated by Hils \& Bender (\cite{HB95}), who
1269: used a method similar to ours, without specifying the criterion for
1270: inspiral.
1271: 
1272: %
1273: \begin{figure}
1274: %\includegraphics[%
1275: %  width=1.0\columnwidth]{f4.EPS}
1276: 
1277: \plotone{f4.eps}
1278: 
1279: \figcaption[FileName]{\label{f:WeMBH}The probability DF of a CO entering the \emph{\LISA}
1280: band with eccentricity $e$ for a MBH of mass $\Mbh\!=\!3\!\times\!10^{6}\,\Mo$.
1281: The histograms represent WDs, NSs and BH (from broad to narrow). Note
1282: that the maximal possible eccentricity for which $P\!<\! P_{L}$ is
1283: $e_{\textrm{max}}\!=\!0.81$. See table (\ref{t:stars}) for the cusp
1284: parameters.}
1285: \end{figure}
1286: 
1287: 
1288: 
1289: \subsection{Intermediate mass black holes in stellar clusters}
1290: 
1291: Unlike MBHs with masses $\Mbh\!\gtrsim\!10^{6}\,\Mo$ , there is little
1292: firm observational evidence at this time for the existence of IBHs
1293: with masses $\Mbh\!\sim\!10^{3}\,\Mo$ (see review by Miller \& Colbert
1294: \cite{MC04}). However, there are plausible arguments arguing in favor
1295: of their existence. From a theoretical point of view, these objects
1296: are thought to form naturally, such as in population III remnants
1297: (Madau \& Rees \cite{MR01}), or in a runaway merger of young stars
1298: in dense stellar clusters (Portegies Zwart \& McMillan \cite{PZM01};
1299: Portegies Zwart et al. \cite{PZea04}). From an observational perspective,
1300: IBHs may power some of the ultraluminous X-ray sources (e.g. Miller,
1301: Fabian, \& Miller \cite{MFM04}), for example by tidal capture of
1302: a main sequence companion star (Hopman et al. \cite{HPZA04}).
1303: 
1304: For the purpose of estimating the orbital parameters of GW emitting
1305: stars, we assume that IBHs lie at the center of dense stellar clusters
1306: (see model parameters listed in table \ref{t:stars}). White dwarfs are
1307: disrupted by an IBH before entering the \emph{\LISA} band, so that
1308: only the most compact sources, neutron stars and stellar mass BHs can
1309: emit GW in the \LISA\ frequency band. The same values for the stellar
1310: population fractions $f_s$ where taken here as for MBHs. We note that
1311: this is not necessarily the case. For example, N-body simulations
1312: indicate that a large fraction of BHs may be ejected in an early stage
1313: of the cluster's life if a massive stellar object forms a tight binary
1314: with the IBH (Baumgardt et el. \cite{Baum04b}).
1315: 
1316: Figure (\ref{f:GWfrac}) shows the inspiral probability functions
1317: $S_{s}(a)$, and Fig. (\ref{f:IBHecc}) shows the eccentricity histograms
1318: of stars on orbits with $P=P_{L}$. The maximum eccentricity still
1319: observable by \LISA\ is nearly unity, and the IBH case shows even
1320: more clearly the strong tendency towards large eccentricities. In
1321: general, this effect becomes more prominent for lighter BHs. The high
1322: eccentricity makes the GW signal highly non-monochromatic. The star
1323: spends most time at apo-apse, emitting a relatively weak, low frequency
1324: signal. At periapse short pulses of high frequency GW is emitted.
1325: For IBHs of $\Mbh\!\sim\!10^{3}\,\Mo$, the frequency $\nu_{p}$ of
1326: these short bursts at periapse is of the order $\nu_{p}\lesssim100$
1327: Hz. This is too high to be measurable by \LISA\ , but may be measurable
1328: by ground based detectors such as \textit{LIGO} or \textit{VIRGO}.
1329: 
1330: %
1331: \begin{figure}
1332: %\includegraphics[%
1333: %  clip,
1334: %  width=1.0\columnwidth,
1335: %  keepaspectratio]{f5.EPS}
1336: 
1337: \plotone{f5.eps}
1338: 
1339: \figcaption[FileName]{\label{f:IBHecc}The probability DF of a compact remnant entering
1340: the \emph{\LISA} band with eccentricity $e$ for an IBH of mass $\Mbh\!=\!10^{3}\,\Mo$.
1341: Only NSs (broad) and BHs (narrow) are considered; WDs are probably
1342: disrupted by the tidal field. For an IBH, the maximal possible eccentricity
1343: for which $P\!<\! P_{L}$ is nearly unity, $e_{\textrm{max}}\!=\!0.998$,
1344: all inspiraling stars are likely to have eccentricities close to the
1345: maximum value. See table (\ref{t:stars}) for the cusp parameters.}
1346: \end{figure}
1347: 
1348: 
1349: As anticipated (Eq. \ref{e:massdependence}), the rate of inspiraling
1350: stars is comparable to that of a MBH (Table \ref{t:stars}). However,
1351: due to the extremely high eccentricities and small semi-major axes
1352: of the orbits, the GW emission is very efficient and they spiral in
1353: on a very short time scale (on the order of a year).
1354: 
1355: 
1356: \section{Summary and discussion}
1357: 
1358: \label{s:sum}
1359: 
1360: Stars spiraling into MBH due to the emission of GW are an important
1361: potential source for future GW detectors, such as \LISA. The detection
1362: of the signal against the noise will be challenging, and requires
1363: pre-calculated wave-forms.  The waveforms depend on the orbital
1364: parameters of the inspiraling stars. Our main goal in this study was
1365: to derive the distribution of eccentricities of inspiraling stars. The
1366: orbital statistics of such stars reflect a competition between orbital
1367: decay through dissipation by the emission of GW, and scattering, which
1368: deflects stars into eccentric orbits but can also deflect them back to
1369: wider orbits or straight into the MBH.
1370: 
1371: Inspiral is a slow process, and unless the stars start close enough to
1372: the MBH, they will be scattered off their orbit. We identified a
1373: critical length scale $a_{c}$ which demarcates the volume from which
1374: inspiral is possible. We obtained an analytical expression for the
1375: inspiral rate and showed that it is of the order of a few per Gyr per
1376: galaxy, much smaller than the rate for direct capture (Alexander \&
1377: Hopman \cite{AH03}), that it is nearly independent of the relaxation
1378: time for typical stellar cusps, and that it grows slowly with
1379: decreasing MBH mass (assuming the $\Mbh-\sigma$ relation). Throughout
1380: we assumed a single powerlaw DF. Generalization to different profiles
1381: is straightforward. Qualitatively, the inspiral rate is determined by
1382: the stellar density near $a_c$, and is not very sensitive to the exact
1383: profile far away at radii much smaller or much larger than $a_c$. The
1384: rate of GW events depends on the number of COs inside $a_{c}$ (which
1385: is much smaller than the MBH radius of influence), and so
1386: mass-segregation can play an important role in enhancing the event
1387: rate by leading to a centrally concentrated distribution of COs.
1388: 
1389: We obtained a relatively low rate. One important reason for this is
1390: our stringent criterion for inspiral. This may also explain why our
1391: results deviate from Hils \& Bender \cite{HB95}, although they do not
1392: specify their precise criterion for inspiral. Comparison with other
1393: works are more complicated. Possible differences may stem from
1394: different normalizations of the central density, different CO
1395: fractions, different criteria for inspiral, and the way mass
1396: segregation is treated. An essential step in the future will be to
1397: analyze inspiral processes by N-body simulations, which are feasible
1398: already for small systems (Baumgardt et al. \cite{Baum04a},
1399: \cite{Baum04b}; Preto et al. \cite{P04}).
1400: 
1401: 
1402: The detection rate depends on the inspiral rates, but also on the mass,
1403: relaxation time, the orbital parameters (especially the eccentricity)
1404: and the details of the detection algorithm (Barack \& Cutler
1405: \cite{BC04a}). Here we provide a simple recipe to estimate the number of
1406: detectable sources.  We assume that the various dependencies above can
1407: be expressed by an effective strain $\hat{h}$.
1408: 
1409: The strain of GW resulting from a star of mass $\Ms=m\Mo$ orbiting a
1410: MBH of mass $\Mbh=10^6\Mo M_6\gg\Ms$ at a distance $d={\rm
1411: Gpc}\,d_{\rm Gpc}$, on a circular orbit of orbital frequency $\nu=
1412: 10^{-3}{\rm s^{-1}}\nu_{-3}$, is given by
1413: 
1414: \begin{equation}\label{eq:strain}
1415: h = 1.7\times 10^{-23}\nu_{-3}^{2/3}d_{\rm Gpc}^{-1}M_6^{2/3}m.
1416: \end{equation}
1417: (e.g., Sigurdsson \& Rees \cite{SR97}).
1418: 
1419: The cosmic rate of {\it LISA} events is given by
1420: 
1421: \begin{equation}\label{eq:cosmrate}
1422: \Gamma_{\rm tot} = 
1423: \int_{M_{\rm min}}^{M_{\rm max}} d\Mbh {d\nbh\over d\Mbh} \Gamma_i(\Mbh) V(\Mbh),
1424: \end{equation}
1425: where $d\nbh/d\Mbh$ is the number of MBHs per unit mass, per unit
1426: volume, and $ V(\Mbh)$ is the volume in which stars can be observed by
1427: {\it LISA}. Aller \& Richstone (\cite{AR02}) used the $\Mbh-\sigma$
1428: relation to ``weigh'' the MBHs. The spectrum is roughly approximated
1429: by
1430: 
1431: \begin{equation}
1432: {d\nbh\over d\Mbh} = 10^7\left({1\over10^6\Mo}\right) M_6^{-1}\, {\rm Gpc^{-3}}.
1433: \end{equation}
1434: 
1435: The LISA sensitivity curve at $\nu=10^{-3}$ is $h\sim10^{-23}$ for one
1436: year of observation with signal to noise ratio S/N=1 (see e.g.  {\it
1437: http://www.srl.caltech.edu/lisa}). We adopt this value as a
1438: representative detection threshold for $\hat{h}$. If the effective
1439: strain has to be be at least $10^{-23}\hat{h}_{-23}$ for the star to
1440: be observable, then, for a Euclidean Universe,
1441: 
1442: \begin{equation}
1443: V(\Mbh) = 20.6\,{\rm Gpc^{3}}\,\hat{h}_{-23}^{-3}\nu_{-3}^2m^3M_6^2.
1444: \end{equation}
1445: 
1446: 
1447: Finally, the rate per MBH is 
1448: \begin{equation}
1449: \Gamma_i(\Mbh) =
1450: \Gamma_i(3\times10^6\Mo)(3M_6)^{-1/4},
1451: \end{equation}
1452: where we used the expression for the dependence of the inspiral rate
1453: on the MBH mass, equation (\ref{eq:massdependence}), with $\beta=4$.
1454: The expression can be calibrated with the MC results for a $3\times
1455: 10^6 \Mo $ MBH.
1456: 
1457: Integrating (\ref{eq:cosmrate}) from $M_6 = 0.5$ to  $M_6=5$ gives
1458: 
1459: \begin{equation}
1460: \Gamma_{\rm tot} = 
1461: 1.5\,\peryr\,\left[{\Gamma_i(3\times10^6\Mo)\over {\rm Gyr^{-1}}}\right]m^3\hat{h}_{-23}^{-3}\nu_{-3}^2.
1462: \end{equation}
1463: 
1464: The number of sources which {\it LISA} can observe at any moment is
1465: estimated by $\mathcal{N} = \Gamma_{\rm tot}\times t_L(\bar{e}, a_L)$,
1466: where $t_L$ is the time a star with eccentricity $\bar{e}$ spends in
1467: the {\it LISA} band before being swallowed; here $\bar{e}$ is the
1468: average eccentricity of stars entering the {\it LISA} band.
1469: 
1470: For example, our calculations for WD inspiral indicate that
1471: $\Gamma_i(3\times10^6\Mo)=7.8\,{\rm Gyr^{-1}}$. For a circular orbit
1472: with a period of $P=10^3\,{\rm s}$ and $M_6=1$, the inspiral time is
1473: $t_L=54\,{\rm yr}$, in which case the number of detectable sources
1474: would be $\mathcal{N} \sim\,130\hat{h}_{-23}^{-3}\nu_{-3}^2$.
1475: 
1476: We would like to emphasize that this estimate is to be treated with
1477: caution. The number of detectable sources depends strongly on the
1478: assumptions. In particular, GWs from eccentric sources are not
1479: monochromatic and may be harder to detect. The analysis in this paper
1480: can be used for a more detailed analysis of the number of detectable
1481: {\it LISA} sources.
1482: 
1483: We used three complementary methods to model the inspiral process, MC
1484: simulations, numerical solution of the diffusion/dissipation equation
1485: and an analytic effective orbit model. We followed the evolution of
1486: the orbital properties of the inspiraling stars to the point where
1487: they decoupled from the scattering. All three methods were in
1488: excellent agreement. We find that the distribution of orbital angular
1489: momenta is strongly peaked near $J\!\gtrsim\! J_{lc}$, with the
1490: detailed form of the distribution depending somewhat on the parameters
1491: of the stellar system. We demonstrated that estimates that
1492: {}``freeze'' the scattering prematurely may lead to erroneously high
1493: rates by counting stars that actually plunge into the MBH. We then
1494: used the most versatile method, the MC simulations, to continue
1495: evolving the orbits in a GR Schwarzschild potential, taking into
1496: account energy and angular momentum loss due to GW emission (and
1497: residual perturbations by scattering).  We derived the distribution of
1498: eccentricities of the inspiraling stars as they enter the detection
1499: window (orbital period of $P_{L}\!\lesssim\!10^{4}\,\mathrm{s}$ for
1500: \LISA).
1501: 
1502: Our main result is that the eccentricities of stars entering the \LISA\ 
1503: band are strongly skewed toward high values.
1504: 
1505: 
1506: 
1507: \acknowledgements{TA is supported by ISF grant 295/02-1, Minerva grant 8484, and a
1508: New Faculty grant by Sir H. Djangoly, CBE, of London, UK.}
1509: 
1510: \begin{thebibliography}{2004b}
1511: \bibitem[1999]{TA99}Alexander, T., 1999, \apj, 520, 137 
1512: \bibitem[2003]{AM03}Alexander, T., \& Morris, M., 2003, \apj, 590, L25 
1513: \bibitem[2003]{AH03}Alexander, T., \& Hopman, C., 2003, \apj, 590, L29 
1514: \bibitem[2002]{AR02}Aller, M. C., \& Richstone, D., 2002, \aj, 124, 3035 
1515: \bibitem[1976]{BW76}Bahcall, J. N., \& Wolf, R. A., 1976, \apj, 209, 214 
1516: \bibitem[1977]{BW77}Bahcall, J. N., \& Wolf, R. A., 1977, \apj, 216, 883 
1517: \bibitem[2003]{BC04a}Barack, L., \& Cutler, C., 2004, PRD, 69, 082005 
1518: \bibitem[2004]{BC04b}Barack, L., \& Cutler, C., 2004, gr-qc/0409010 
1519: \bibitem[2004a]{Baum04a}Baumgardt, H., Makino, J., \& Ebisuzaki, T., 2004a, \apj, 613, 1133
1520: \bibitem[2004b]{Baum04b}Baumgardt, H., Makino, J., \& Ebisuzaki, T., 2004b,\apj, 613, 1143
1521: \bibitem[1987]{BT87}Binney, J. \& Tremaine, S., 1987, Galactic Dynamics (Princeton: Princeton
1522: Univ. Press) 
1523: \bibitem[1978]{CK78}Cohn, H., \& Kulsrud, R. M. 1978, \apj, 226, 1087 
1524: \bibitem[1994]{CKP94}Cutler, C., Kennefick, D., \& Poisson, E., 1994, PRD, 50, 6
1525: \bibitem[2000]{FM00}Ferrarese, L., \& Merritt, D., 2000, \apj, 539, L9 
1526: \bibitem[1999]{Fig99}Figer, D. F., Kim, S. S., Morris, M., Serabyn, E., Rich, R. M., \&
1527: McLean, I. S. 1999, \apj, 525, 750
1528: \bibitem[1976]{FR76}Frank, J., \& Rees, M. J., 1976, \mnras, 176, 633 
1529: \bibitem[2001]{FB01}Freitag, M., \& Benz, W, 2001, A\&A, 375, 711 
1530: \bibitem[2001]{FR01}Freitag, M., 2001, Class. Quantum Grav., 18, 4033 
1531: \bibitem[2003]{FR03}Freitag, M., 2003, \apj, 583, L21 
1532: \bibitem[2002]{Fre02}Freitag, M. \& Benz, W., 2002, A\&A, 394, 345
1533: \bibitem[2004]{Gai04}Gair, J. R., Barack, L., Creighton, T., Cutler, C., Larson, S., L.,
1534: Phinney, E. S., Vallisneri, M., 2004, gr-qc/0405137 
1535: \bibitem[2000]{Geb00}Gebhardt, K., et al., 2000, \apj, 539, L13 
1536: \bibitem[2003]{Geb03}Gebhardt, K., et al. 2003, \apj, 583, 92 
1537: \bibitem[2003]{Genzelea03}Genzel, R. et al., 2003, \apj, 594, 812 
1538: \bibitem[2003]{GS03}Genzel, R., Sch\"{o}del, R., Ott, T., Eckart, A., Alexander, T.,
1539: Lacombe, F., Rouan, D., \& Aschenbach, B., 2003, \nat, 425, 934 
1540: \bibitem[2002]{GHK02}Glampedakis, K., Hughes, S. A., \& Kennefick, D., 2002, \prd, 66,
1541: 064005 
1542: \bibitem[1995]{HB95}Hils, D., \& Bender, P. L., 1995, \apj, 445, L7 
1543: \bibitem[2004]{HPZA04}Hopman, C., Portegies Zwart, S.F., \& Alexander, T., 2004, \apj,
1544: 604, L101 
1545: \bibitem[2002]{IV02}Ivanov, P. B., 2002, \mnras, 336, 373, 2002 
1546: \bibitem[1977]{LS77}Lightman, A. P., \& Shapiro, S. L., 1977, \apj, 211, 244 
1547: \bibitem[2001]{MR01}Madau, P., \& Rees, M. J., 2001, \apj 551, L27 
1548: \bibitem[1999]{MT99}Magorrian, J., \& Tremaine, S., 1999, \mnras, 309, 447 
1549: \bibitem[2004]{MC04}Miller, M. C., \& Colbert, J. M., 2004, Int.J.Mod.Phys. D13 1-64 
1550: \bibitem[2004]{MFM04}Miller, J. M., Fabian, A., C., \& Miller, M. C., 2004, 614, \apj,
1551: L117 
1552: \bibitem[2000]{MG00}Miralda-Escud\'{e}, J., \& Gould, A., 2000, \apj, 545, 847 
1553: \bibitem[2004]{MO04}Mouawad, N., Eckart, A., Pfalzner, S.,
1554: Sch\"{o}del, R., Moultaka, J., Spurzem, R., 2004, Astronomische
1555: Nachrichten, Vol. 326, 2, 83-95
1556: \bibitem[1972]{P72}Peebles, P. J. E., 1972, \apj, 178, 371 
1557: \bibitem[1964]{Pe64}Peters, P. C. 1964, Phys. Rev., 136, 1224 
1558: \bibitem[2001]{PPSLR01}Pierro, V., Pinto, I. M., Spallicci, A. D., Laserra, E., Recano, F.,
1559: 2001, \mnras, 325, 358 
1560: \bibitem[2001]{PZM01}Portegies Zwart, S. F., McMillan, S. L. W. 2002, \apj, 576, 99 
1561: \bibitem[2004]{PZea04}Portegies Zwart, S. F., Baumgardt, H., Makino, J., McMillan, S. L.,
1562: \& Hut, P., \nat, 428, 724 
1563: \bibitem[1992]{PTVF77}Press, W. H., Teukolsky, S. A., Vetterling, W. T., \& Flannery, B.
1564: P. 1992, (Cambridge: University Press), 2nd ed.
1565: 
1566: \bibitem[2004]{P04}Preto, M., Merritt, D., Spurzem, R., 2004, \apj, 613, L109
1567: \bibitem[1995]{QHS95}Quinlan, G. D., Hernquist, L., \& Sigurdsson, S., 1995, \apj, 440, 554
1568: \bibitem[2003]{S03}Sigurdsson, S., astro-ph/0304251 
1569: \bibitem[1997]{SR97}Sigurdsson, S., and Rees, M. J., 1997, \mnras 284, 318 
1570: \bibitem[1999]{SU99}Syer, D., \& Ulmer, A., 1999, \mnras, 306, 35 
1571: \bibitem[2002]{Tr02}Tremaine, S., et al., 2002, \apj, 574, 740 
1572: \bibitem[2004]{WM04}Wang, J., \& Merritt, D., 2004, \apj, 600, 149
1573: \bibitem[2005]{WG05}Wen \& Gair, 2005, pre-print: gr-qc/0502100
1574: \bibitem[1980]{You80}Young, P., 1980, \apj, 242, 1232
1575: \bibitem[2002]{ZHR02}Zhao, H.-S., Haehnelt, M. G., \& Rees, M. J., 2002, New Astron., 7,
1576: 385\end{thebibliography}
1577: 
1578: \end{document}
1579: