1: \documentclass[useAMS,usenatbib, usegraphicx]{mn2e}
2:
3: \topmargin -0.6in
4:
5: \usepackage{epsfig}
6: \usepackage{bm}
7: %\usepackage{lscape}
8: %\usepackage{graphicx}
9: %\usepackage{color}
10: \usepackage{amsmath}
11: \usepackage{amssymb}
12: %\usepackage{amsfonts}
13: %\usepackage{amstext}
14: %\usepackage{amsbsy}
15: \usepackage{natbib}
16: %\loadbold
17: \bibliographystyle{mn2e}
18: %\newcommand{\tbfrac}[2]{\genfrac{}{}{0pt}{0}{#1\strut}{#2\strut}}
19: \newcommand{\kms}{\ensuremath{{\rm km\,s}^{-1}}}
20: \newcommand{\msun}{\ensuremath{{\rm M}_{\odot}}}
21: \newcommand{\rsun}{\ensuremath{{\rm R}_{\odot}}}
22: \newcommand{\mbh}{\ensuremath{M_{\rm BH}}}
23: \newcommand{\mm}{\ensuremath{M_m}}
24: \newcommand{\mM}{\ensuremath{M_M}}
25: \newcommand{\yr}{\ensuremath{\rm yr}}
26: \newcommand{\myr}{\ensuremath{\rm Myr}}
27: \newcommand{\msmbh}{\ensuremath{M_{\rm SMBH}}}
28: \newcommand{\infinity}{{\infty}}
29: \newcommand{\apj}{ApJ}
30: \newcommand{\apjl}{ApJ}
31: \newcommand{\mnras}{MNRAS}
32: \newcommand{\aj}{AJ}
33: \newcommand{\apjs}{ApJS}
34: \newcommand{\nat}{Nat}
35: \newcommand{\pasj}{PASJ}
36: \newcommand{\aap}{A\&A}
37: \newcommand{\letter}{{paper}}
38: \newcommand{\sag}{Sgr~A*}
39: \newcommand{\rmd}{{\rm d}}
40: \newcommand{\physrep}{Physics Reports}
41: \newcommand{\aaps}{A\&AS}
42: \newcommand{\araa}{ARA\&A}
43: \newcommand{\mtot}{\ensuremath{M_{\rm tot}}}
44:
45:
46:
47: \newcommand\lsim{\mathrel{\rlap{\lower4pt\hbox{\hskip1pt$\sim$}}
48: \raise1pt\hbox{$<$}}}
49: \newcommand\gsim{\mathrel{\rlap{\lower4pt\hbox{\hskip1pt$\sim$}}
50: \raise1pt\hbox{$>$}}}
51: \newcommand\propsim{\mathrel{\rlap{\lower4pt\hbox{\hskip1pt$\sim$}}
52: \raise1pt\hbox{$\propto$}}}
53: \newcommand{\Si}{\mathrm{Si}}
54: \newcommand{\D}{\mathrm{d}}
55: \newcommand{\LISA}{\textit{LISA\,}}
56: \newcommand{\JWST}{\textit{JWST\,}}
57: \newcommand{\JDEM}{\textit{JDEM\,}}
58: \newcommand{\EM}{\mathrm{em}}
59: \newcommand{\fast}{\mathrm{fast}}
60: \newcommand{\slow}{\mathrm{slow}}
61: \newcommand{\spin}{\mathrm{spin}}
62: \newcommand{\spinang}{SA}
63: \newcommand{\spinmag}{SM}
64: \newcommand{\dL}{d_{\rm L}}
65: \newcommand{\dy}{\,\mathrm{days}}
66: \newcommand{\hr}{\,\mathrm{hr}}
67: \newcommand{\Mpc}{\,\mathrm{Mpc}}
68: \newcommand{\AU}{\,\mathrm{AU}}
69: \newcommand{\mHz}{\,\mathrm{mHz}}
70: \newcommand{\Hz}{\,\mathrm{Hz}}
71: \newcommand{\Mchirp}{\mathcal{M}}
72: \newcommand{\Msun}{\mathrm{M}_{\odot}}
73: \newcommand{\ii}{i}
74: \newcommand{\arun}{2008PhRvD..77f4035A}
75: \newcommand{\baker}{2007PhRvD..75l4024B}
76: \newcommand{\bc}{2004PhRvD..69h2005B}
77: \newcommand{\berti}{2007PhRvD..76j4044B}
78: \newcommand{\blanchet}{2006LRR.....9....4B}
79: \newcommand{\gairof}{2005PhRvD..72h4009G}
80: \newcommand{\gairos}{2006PhRvD..74j9901G}
81: \newcommand{\hindera}{2008PhRvD..77h1502H}
82: \newcommand{\hinderb}{2008arXiv0806.1037H}
83: \newcommand{\pmat}{1963PhRv..131..435P}
84: \newcommand{\pk}{2007CQGra..24...83P}
85: \newcommand{\peters}{1964PhRv..136.1224P}
86: \newcommand{\w}{2008arXiv0802.2520W}
87: \newcommand{\will}{2006LRR.....9....3W}
88: \newcommand{\tu}{1977ApJ...216..610T}
89: \newcommand{\prd}{Phys.\ Rev.\ D}
90: \newcommand{\morris}{1993ApJ...408..496M}
91: \newcommand{\meg}{2000ApJ...545..847M}
92: \newcommand{\new}{}
93:
94: %\begin{document}
95: \title[GWs from BHs in Galactic Nuclei]{Gravitational waves from scattering of stellar-mass black holes in galactic nuclei}
96:
97:
98:
99: \author[O'Leary, Kocsis,\& Loeb]{Ryan M.\
100: O'Leary$^{1}$\thanks{E-mail:roleary@cfa.harvard.edu}, Bence Kocsis$^{1.2}$\thanks{E-mail:bkocsis@ias.edu}, and Abraham
101: Loeb$^{1}$\thanks{E-mail:aloeb@cfa.harvard.edu}\\$^{1}$Harvard-Smithsonian Center for Astrophysics, 60 Garden St., Cambridge, MA 02138, USA\\$^{2}$Institute for Advanced Study, Einstein Drive, BH-151, Princeton, NJ 08540}
102:
103: \begin{document}
104: \maketitle
105:
106: \begin{abstract}
107: Stellar mass black holes (BHs) are expected to segregate and form a
108: steep density cusp around supermassive black holes (SMBHs) in
109: galactic nuclei. We follow the evolution of a multi-mass system of
110: BHs and stars by numerically integrating the Fokker-Planck energy
111: diffusion equations for a variety of BH mass distributions. We find
112: that the BHs ``self-segregate'', and that the rarest, most massive
113: BHs dominate the scattering rate closest to the SMBH ($\lesssim
114: 10^{-1}\,$pc). BH--BH binaries form out of gravitational wave
115: emission during BH encounters. We find that the expected rate of BH
116: coalescence events detectable by Advanced LIGO is {\new $\sim
117: 1-10^2\,$yr$^{-1}$, depending on the initial mass function of
118: stars in galactic nuclei and the mass of the most massive BHs. We
119: find that the actual merger rate is likely $\sim 10$ times larger
120: than this due to the intrinsic scatter of stellar densities in
121: many different galaxies.} The BH binaries that form this way in
122: galactic nuclei have significant eccentricities as they enter the
123: LIGO band ($90\%$ with $e > 0.9$), and are therefore distinguishable
124: from other binaries, which circularise before becoming
125: detectable. We also show that eccentric mergers can be detected to
126: larger distances and greater BH masses than circular mergers, up to
127: $\sim 700\,\Msun$. Future ground-based gravitational wave
128: observatories will be able to constrain both the mass function of
129: BHs and stars in galactic nuclei.
130: \end{abstract}
131:
132: \begin{keywords}
133: galaxies:kinematics and dynamics--galaxies:nuclei--black hole
134: physics--gravitational waves
135: \end{keywords}
136:
137:
138: \section{Introduction}
139: \subsection{Motivation}\label{sec:motivation}
140: The coalescence of stellar mass black hole-black hole (BH-BH) binaries
141: is one of the most anticipated sources of gravitational waves for
142: ground-based interferometers such as LIGO\footnote{http://www.ligo.caltech.edu/} or VIRGO\footnote{http://www.virgo.infn.it/}.
143: Whether formed primordially \citep{2002ApJ...572..407B, 2004ApJ...611.1068B,
144: 2008ApJ...676.1162S}
145: or in dense star clusters \citep{2000ApJ...528L..17P, 2004ApJ...616..221G,
146: 2006ApJ...637..937O, 2006ApJ...640..156G, 2007PhRvD..76f1504O,
147: 2008arXiv0804.2783M}, nearly all of these binaries are expected to be
148: circularised by gravitational wave (GW) emission before they are detectable
149: by ground based observatories \citep[][]{2006ApJ...637..937O,
150: 2006ApJ...640..156G}. In order to be detected with a high signal-to-noise
151: ratio, matched filtering algorithms sift through the LIGO data stream
152: looking for such circular inspirals \citep{2008PhRvD..77f2002A}, but might
153: miss many eccentric events
154: {\new \citep{PhysRevD.60.124008,2007arXiv0712.3199T,2008ApJ...681.1431M}}.
155: In principle, eccentric inspirals are well suited for detection as well, however they
156: have not been expected to be a significant source for terrestrial detectors
157: \citep[except see][]{2003ApJ...598..419W}.
158:
159: In this paper, we propose an important additional source of
160: gravitational wave sources, which preferentially form binaries that
161: are still eccentric as they merge. In galactic nuclei with
162: supermassive black holes (SMBH) of mass $\msmbh < 10^7\,\Msun$,
163: relaxation times are often less than a Hubble time, and can result in
164: the formation of steep density cusp of stars and stellar mass black holes (BHs). Indeed, as many
165: as $\sim 20,000$ BHs are expected to have
166: segregated into the inner $\approx 1\,$pc of the Milky Way
167: \citep[][hereafter HA06]{1993ApJ...408..496M,2000ApJ...545..847M,
168: 2006ApJ...649...91F,2006ApJ...645L.133H}. When two BHs have a close
169: encounter, they can release sufficient amount of energy in GWs to form
170: a tight binary, and merge in less than a few hours. These mergers are
171: almost always eccentric, in contrast to circular inspirals
172: expected in globular clusters. A similar process can occur in the
173: runaway growth of BHs when no SMBH is present that will also result in
174: eccentric mergers; however, such systems are inherently unstable and
175: not long-lived
176: \citep{1987ApJ...321..199Q,1989ApJ...343..725Q,1990ApJ...356..483Q,1993ApJ...418..147L}.
177:
178: The GWs generated during the evolution of these orbits are endowed with a
179: rich structure. For sufficiently small impact parameters, the GWs can be
180: detected with second generation GW instruments already during first passage
181: \citep{2006ApJ...648..411K}. Subsequently, the eccentricity is close to
182: unity for several thousand orbits after the formation of the binary,
183: generating a train of short-duration distinct GW bursts with a continuous
184: frequency spectrum. As the eccentricity gradually decreases, the
185: separation between subsequent bursts decreases, and the waveform becomes
186: continuous in the time-domain and decouples into discrete harmonics in
187: frequency. Interestingly, as we show below, the eccentricity is
188: non-negligible all the way to coalescence. The full waveform,
189: consisting of a correlated set of distinct GW bursts, which evolves into a
190: continuous eccentric inspiral signal, would not be discovered in the GW data
191: by existing data analysis techniques such as GW burst search algorithms,
192: nor by small eccentricity inspiral templates, since the contribution of
193: individual bursts to the GW power is small and the contribution of higher
194: harmonics is significant throughout the evolution
195: \citep[see][for a similar discussion for extreme mass ratio GW bursts during encounters with a supermassive black hole]{2008ApJ...675..604Y}.
196:
197: Eccentric inspirals are more luminous in GWs than circular
198: inspirals and extend to higher frequencies, and so may be detectable to
199: higher redshifts. Furthermore, these events may be capable of detecting
200: intermediate mass black holes using terrestrial instruments
201: \citep{2007PhRvL..99t1102B,2008ApJ...681.1431M}. The evolution of the signal during the
202: initial phase of eccentric inspirals can be described using a small number
203: of parameters, allowing one to construct very efficient and sensitive data
204: analysis algorithms to search for these signals in the LIGO
205: data. Additionally, these signals spend a longer physical time in the LIGO
206: band, possibly reducing the false alarm rate for a fixed number of templates that
207: cover the observational period. The modulation introduced by the Earth's rotation
208: during such an event can be utilised to further improve the measurement
209: accuracy for the source position.
210:
211: The GW waveforms have been studied for eccentric inspirals through several
212: different approaches. Post-Newtonian (PN) waveforms exist up to 1.5PN
213: order beyond the leading order gravitational radiation effects ($\propto v^3$)
214: for general mass ratios with spins
215: \citep{Vasuth:2007gx,2008PhysRevD.77.104005},
216: {\new 2PN without spins \citep[($\propto v^4$, including the effects of
217: eccentricity and radiation reaction]{2004PhRvD..70f4028D} and 3PN
218: including eccentricity but without spins and also neglecting radiation
219: reaction \citep[$\propto v^6$,][and references
220: therein]{2008PhRvD..77f4035A}, 5.5PN for extreme mass ratios without spins
221: \citep[$\propto v^{11}$,][]{1996PThPh..96.1087T}. The evolution of a
222: binary is chaotic at (or beyond) the 2PN approximation if the component
223: BHs have spins \citep{2006PhRvD..74l4027L,2007PhRvD..76l4004W} and also
224: possibly for nonspinning BHs approaching the unstable circular
225: orbit \citep[i.e. exhibiting zoom-whirl
226: orbits,]{2007CQGra..24...83P,2008arXiv0802.2520W}.}
227: Numerical relativity has mostly focused on the circular inspiral for
228: comparable mass BHs \citep[and references therein]{\baker,\berti,2008arXiv0804.4184B}, or extreme
229: mass ratios \citep{2004PhRvD..69d4025M,2007CQGra..24...83P,2007PhRvD..75b4005B}.
230: Recently, however, the waveforms have been evaluated for a handful of equal mass eccentric
231: initial conditions
232: \citep{2007arXiv0710.3823S,2008arXiv0802.2520W,2008arXiv0806.1037H,2008PhRvD..77h1502H}.
233:
234: In this paper, we perform two separate analyses to two different
235: problems. First, in \S~\ref{sec:massseg}, we determine the multi-mass
236: distribution of BHs in galactic nuclei for a variety of mass models.
237: In \S~\ref{sec:GB}, we determine the formation rate of
238: binaries due to gravitational wave capture in a variety of galactic
239: nuclei. In \S~\ref{sec:detection} we describe the general features
240: of the generated GW waveform, calculate the expected signal-to-noise ratio
241: for its detection with second generation terrestrial GW instruments and
242: determine the expected detection rate of such mergers. We also show that
243: the eccentricity distribution of such sources will distinguish binaries
244: formed in the manner outlined in this paper from BH-BH binaries formed
245: dynamically in massive star clusters.
246:
247:
248: \section{Mass Segregation}
249: \label{sec:massseg}
250:
251:
252: \citet[hereafter BW76]{1976ApJ...209..214B} were the first to
253: correctly analyse the relaxation of a stellar population around a
254: central point mass, {\new although their analysis was highly idealised}. They
255: first derived the Fokker-Planck equations for a spherically symmetric
256: distribution of a single-mass population of low mass stars around a
257: massive black hole ($M_* \ll \msmbh$). They found that after about
258: one half of a relaxation timescale the mass density profile of stars
259: reaches a steady-state and forms a power-law cusp with respect to
260: radius, $\propto r^{-\alpha}$, around the massive central object, with
261: $\alpha = 7/4$. In a second paper, they extended their analysis to
262: look at a multi-mass system and included effects of the loss-cone
263: \citep{1976Natur.262..743S} that results from the disruption of stars
264: by the central black hole \citep[][hereafter
265: BW77]{1977ApJ...216..883B}. They found that in a two-mass system the
266: more massive objects segregate from the lower mass stars by forming a
267: steeper power-law density profile than the stars. This is in stark
268: contrast to the evolution of a two-mass cluster of stars without a
269: SMBH, which can eventually lead to the so-called Spitzer-instability
270: \citep{1969ApJ...158L.139S}, in which the high mass stars decouple
271: from the low mass objects entirely.
272:
273: For a population of stars that are sufficiently old ($\ga 100\,\myr$),
274: stellar mass BHs with mass $\sim 10\,\Msun$ are expected to become the most
275: massive objects in the system and will begin to segregate from the main
276: sequence stars due to dynamical friction \citep{\morris}. Looking at our
277: own Galactic Center and using a typical stellar mass function, \citet{\meg}
278: estimated that $\sim 2.5 \times 10^4\,$ BHs should have segregated into the
279: inner pc. Indeed, HA06 followed the work of BW77 by solving the time
280: dependent Fokker-Planck equations and confirmed these results. They found
281: that $\sim 18,000$ BHs of $10\,\Msun$ each, should segregate to form a
282: steep density cusp around the SMBH in our galaxy, \sag, with $\alpha_{\rm
283: BH} \approx 2$, and that the stars and lighter compact objects form a
284: shallower cusp with $\alpha_* \approx 1.4$.
285:
286: \citet[hereafter FAK06]{2006ApJ...649...91F} have so far provided the most
287: robust analysis of the segregation and distribution of a multi-mass system
288: of BHs and stars in the nuclei of galaxies \citep[for a review, see][]{2007astro.ph..3495A}.
289: The authors combined large $N\sim 10^6$ Monte-Carlo Fokker-Planck simulations with a population
290: synthesis code \citep{2002ApJ...572..407B} to simulate the segregation and
291: subsequent formation of a density cusp for an old population of stars.
292: Their simulations were in remarkably good agreement with previous analyses
293: \citep[e.g.][HA06]{\morris, \meg}. In their simulations of a
294: Milky-Way--like galactic nucleus they found a similar overall number of
295: stellar mass BHs ($\sim 20,000$)
296: within 1$\,$pc of a $3.5\times10^6\Msun$ SMBH. However, they found the BHs
297: follow a slightly shallower density profile than HA06, with $\alpha_{\rm
298: BH} \approx 1.8$. Despite the level of detail in the simulations of FAK06,
299: the simulations still have their limitations. Because of the large range
300: in scales needed to follow the mass segregation, and subsequently the large
301: number of particles in that volume, each star cannot be modelled by a
302: single particle. The densest, most interesting regions of their
303: simulations still suffer from small number statistics, and can't reveal the
304: precise expectations for the mass distribution of stars and BHs. In order
305: to calculate the rate of gravitational wave capture events one needs to
306: apply a high-resolution method (like HA06) to multi-mass distribution
307: of BHs for calculating the rates and detectability of the events.
308:
309: Although the BHs are expected to dominate relaxation and dynamical
310: encounters in the inner $0.1\,$pc around \sag, their presence has so
311: far eluded detection. X-ray observations of our own galaxy and others
312: may provide the first window to this interesting stellar population.
313: In our own galaxy, \citet{2005ApJ...622L.113M} have found an
314: overabundance of X-ray sources in the vicinity of \sag, which seems to
315: follow the underlying distribution of stars
316: \citep[][]{2006ApJS..165..173M}. For large distances, the X-ray
317: sources appear to follow the BH distribution as well (FAK06). In
318: other galaxies, many BHs may interact with the dense accretion disks
319: around SMBHs, and in some circumstances be intrinsically more luminous
320: than the SMBH \citep{2007MNRAS.377..897D, 2007MNRAS.377.1647N}.
321:
322: The existence of a nuclear population of BHs may also be revealed
323: through dynamical encounters with observable stars. Indeed,
324: \citet{2005ApJ...622..878W} have shown that a decade of observations
325: with the next generation of $30\,$m class telescopes may detect the
326: scattering of the BHs with the so-called 'S-stars' that are within
327: $\approx 0.04\,$pc of \sag. Microlensing by the BHs is expected to
328: produce a much weaker signature
329: \citep{2000ApJ...545..847M,2001ApJ...551..223A, 2001ApJ...563..793C}.
330: Strong encounters between the BHs and the same population of S-stars
331: may dynamically eject the stars with high enough velocity to escape
332: the potential of the galaxy \citep{2008MNRAS.383...86O} and populate
333: the halo with the observed hypervelocity stars
334: \citep{2005ApJ...622L..33B,2006ApJ...640L..35B}. Strong encounters
335: are also expected to tidally spin up long-lived stars, and so the BH
336: population may be inferred through spectroscopic measurements of the
337: spin of lower-mass stars \citep{2001ApJ...549..948A}. {\new
338: Alternatively, one may infer the presence of the BHs by looking at
339: density distribution of old pulsars \citep{2002ApJ...571..320C}.}
340:
341: Finally, future GW observations with the space based GW interferometer
342: {\em LISA}\footnote{http://www.lisa-science.org/} or the Earth-based
343: second generation GW instruments are expected to detect the inspiral
344: of such BHs into SMBHs
345: \citep{2004CQGra..21S1595G,2006ApJ...649L..25R,2007MNRAS.378..129H,2008ApJ...675..604Y}
346: or IMBHs \citep{2007PhRvL..99t1102B,2008ApJ...681.1431M},
347: respectively, several times per year. Given the test particle limit
348: of the BH, such encounters are expected to be a stringent test of
349: General Relativity in the strong-field regime
350: \citep{2004PhRvD..69l4022C,2006PhRvD..74b4006A}.
351:
352:
353: An interesting method for detecting a dense population of BHs is
354: through the GWs they produce as they scatter on each other and get
355: captured into a tight binary. \citet{2006ApJ...648..411K} estimated
356: the expected detection rates of the first hyperbolic passage between
357: such stellar mass black holes in globular clusters, and found rates to
358: be typically less than $0.1{\,\rm yr}^{-1}$ for second generation GW
359: instruments. Here, we consider the detectability of the much stronger
360: GWs that follow once the eccentric binary has formed in galactic
361: nuclei in this way and evolves toward coalescence. We explore this
362: new method in \S~\ref{sec:GB} and~\ref{sec:detection}. However, in
363: order to completely determine this rate we need to know how a
364: multi-mass distribution of BHs relaxes around a SMBH in galactic
365: nuclei. In the remainder of this section we calculate the
366: steady-state distribution of BHs for a variety of models.
367:
368: \subsection{Fokker-Planck Equations}
369: In our calculations, we follow the analysis of BW77, who derived the
370: multi-mass Fokker-Planck equations for a spherically symmetric and
371: isotropic distribution of stars around a SMBH, $f_M(r,v) = f_M(E)$,
372: where $E=(1/2) M v^2 - G M \msmbh/r$ is the mechanical energy of bound
373: orbits in the galactic cusp that are outside the loss cone (see below).
374: Much of our analysis follows that of HA06, whose
375: terminology we use throughout our calculations. The main difference
376: in our work is that we look at a multi-mass distribution of BHs with
377: slightly different initial conditions.
378:
379: The SMBH determines the motion and dynamics of stars within its radius
380: of influence, $r_i = G \msmbh / \sigma_*^2$, where $\msmbh$ is the
381: mass of the SMBH, and $\sigma_*^2$ is the one-dimensional velocity
382: dispersion of the underlying stellar population in the nucleus near
383: the SMBH. For our calculations, we define a dimensionless time
384: variable $\tau = t/t_r$, where $t_r$ is the relaxation timescale at
385: the radius of influence,
386: \begin{equation}
387: \label{eq:relaxtime} t_r = \frac{3(2\pi \beta_* /M_*)^{3/2}}{32 \pi^2 G^2
388: M_*^2 n_* \ln{\Lambda}},
389: \end{equation}
390: where $n_*$ is the number density of stars at $r_i$ for the dominate
391: population of stars of mass $M_*$, $\beta_* = M_* \sigma_*^2$, and $\ln
392: \Lambda \approx \ln (\msmbh/M_*)$ is the Coulomb logarithm. We also define
393: a dimensionless unit of energy, $x = -(M_*/M)(E/\beta_*)$, and a
394: dimensionless distribution function $g_M(x) = [(2 \pi \beta/M_*)^{3/2}
395: n_*^{-1}]f_M(E)$. In these units, the Fokker-Planck equation is (BW77,HA06)
396: \begin{equation}
397: \label{eq:fokkerplanck}
398: \frac{\partial g_M(x,\tau)}{\partial \tau} = -x^{5/2}
399: \frac{\partial}{\partial x} Q_M(x) - R_M(x),
400: \end{equation}
401: where
402: \begin{eqnarray}
403: \label{eq:flowrate}
404: Q_M(x) = \sum_{m} \frac{M}{M_*} \frac{m}{M_*} \int_{-\infinity}^{x_D}
405: \rmd x' [{\max} (x,x')]^{-3/2} \nonumber \\
406: \times \left(g_M(x) \frac{\partial
407: g_{m}(x')}{\partial x'} - \frac{m}{M}g_{m}(x')\frac{\partial
408: g_M(x)}{\partial x}\right).
409: \end{eqnarray}
410: is the rate stars of mass $M$ flow to energies larger than $x$, and
411: $R_M(x)$ is the rate stars are destroyed by the SMBH. We are
412: interested in the distribution of BHs in the inner pc and therefore
413: look only at the empty-loss cone regime. Thus, the angular momentum
414: averaged destruction rate of stars of mass $M$ with energy $x$ is (HA06)
415: \begin{equation}
416: \label{eq:losscone}
417: R_M(x) = \frac{g_M(x)}{\tau_r(x) \ln[J_c(x)/J_{\rm LC}]},
418: \end{equation}
419: where $\tau_r$ is the approximate
420: local relaxation time at radius $r = r_i/(2 x)$,
421: \begin{equation}
422: \label{eq:taur}
423: \tau_r = \frac{M_*^2}{\sum_{M} g_{M}(x) M^2},
424: \end{equation}
425: $J_c(x) = G \msmbh \sigma_*^{-1} (2 x)^{-1/2}$
426: is the angular momentum of a
427: circular orbit at a radius $r = r_i / (2 x)$, and $J_{\rm LC}$ is the
428: maximum angular momentum for the object to be destroyed or consumed by
429: the SMBH. For stars
430: $J_{\rm LC} = G \msmbh \sigma_*^{-1} (2 (x + x_{\rm td}))^{1/2} x_{\rm td}^{-1}$
431: corresponds to a star coming within tidal
432: disruption radius of the SMBH, where
433: $x_{\rm td} = (\msmbh/M_*)^{-1/3} r_i/R_*$
434: and $R_*$ is the radius of the star. BHs plunge into
435: the SMBH when $J < J_{\rm LC} = 4 G \msmbh/c$.
436:
437: We numerically integrate Eq.~(\ref{eq:fokkerplanck}) with our
438: initial conditions given in \S~\ref{initconditions} until $g_M(x)$
439: reaches a steady state for all $M$.
440: For all of our runs this occurs before $\tau = 1$.
441: A computationally straightforward simple iterative way to generate the
442: steady state solution for Eq.~(\ref{eq:fokkerplanck})
443: satisfying $\frac{\partial g_M}{\partial \tau}\equiv 0$
444: is to follow the time evolution of the distribution from the initial
445: conditions.
446:
447: After the integration is complete, we calculate the number density of
448: stars at radius $r$,
449: \begin{equation}
450: \label{eq:numdens}
451: n_M(r) = 2 \pi^{-3/2} n_* \int_{-\infinity}^{\phi(r)}\rmd x g_M(x) (\phi(r)-x)^{1/2},
452: \end{equation}
453: where {\new $\phi(r)$} is the dimensionless specific
454: potential. Until now, we have assumed that the stars were on entirely
455: Keplerian orbits around the SMBH (i.e. that $\phi(r) = r_i/r$). This
456: assumption was necessary in order to simplify the Fokker-Planck
457: equations to Eq.~(\ref{eq:fokkerplanck}) (BW76). However, for our
458: calculations in \S~\ref{sec:GB}, it is important that the BH density
459: goes to zero as $r$ approaches infinity. {\new We therefore calculate
460: $\phi(r)$ and $n_M(r)$ iteratively after calculating the steady-state
461: distribution functions $g_M(x)$. We do this by sequentially solving for $n_M(r)$
462: (Eq.~\ref{eq:numdens}) and
463: \begin{equation}
464: \label{eq:phi}
465: \phi(r) = \frac{r_i}{ r} - \int_0^r \frac{M(<r) r_i}{
466: \msmbh r} \rmd r
467: \end{equation}
468: where $M(<r) = \sum_M \int_0^r 4 \pi r^2 n_M(r) \rmd r$ is the total
469: mass of stars interior to $r$, thus accounting for the stars'
470: contribution to the potential as well as the SMBH's contribution. We
471: find that we converge to a single solution for $\phi(r)$ and $n_M(r)$
472: before about 6 iterations. This solution to the number density and
473: potential is fully consistent with the distribution functions
474: $g_M(x)$. However, Equations~\ref{eq:fokkerplanck} \&
475: \ref{eq:flowrate} implicitly assume that, as $r \rightarrow \infty$,
476: the potential approaches zero and the number density approaches a
477: constant. This assumption is approximately followed as long as the
478: core radius of the isothermal sphere $r_c = [9\sigma_*^2/(4\pi G
479: \rho_0)]^{1/2}$ \citep{1987gady.book.....B} is larger than the radius
480: of influence of the SMBH.} We also note that the number density is the
481: statistical average. In cases of the small numbers of BHs, we assume
482: that values we determine will match the average over many
483: galaxies. However, we caution that the dynamics of such systems will
484: likely behave differently than assumed here.
485:
486:
487: \subsection{Initial Conditions and BH Mass Distributions}
488: \label{initconditions}
489: The initial conditions of our models are described entirely by the
490: distribution functions of unbound stars $g_M(x<0)$.
491: Sufficiently far from
492: the radius of influence of the SMBH the population of visible stars appear
493: to be similar to an isothermal sphere, with a constant velocity dispersion
494: and a number density distribution that scales as
495: $\propto r^{-1.8}$
496: \citep{2003ApJ...594..812G,2007AA...469..125S}.
497: We therefore set $g_*(x) = \exp(x)$ for $x<0$, {\new which corresponds to an isothermal sphere of stars}.
498: We look at only one population of stars of mass $M_* =
499: 1\,\Msun$. This is justified by previous analyses of mass segregation that
500: show that the underlying stellar population, being considerably less
501: massive than BHs, tends to follow the same distribution function
502: independent of mass where the BHs dominate the relaxation (HA06, FAK06).
503: We confirm this conclusion in \S~\ref{sec:massseg:results}. Although there
504: is a significant fraction of more massive stars in nuclei, their lifetimes
505: are expected to be less than the relaxation timescale at most $r$ and {\new are in such small number} that
506: they should not contribute to the dynamics significantly.
507:
508:
509: We assume that the BHs initially follow the distribution of the stars, and
510: amount to a total fraction, $C_{\rm BH} = \sum_M C_M$, of the number
511: density. BH natal kick velocities are expected to be much less than
512: the velocity dispersions of the systems we are interested in
513: \citep{1996ApJ...473L..25W,2005ApJ...625..324W}, and so we expect the
514: BHs' velocity dispersion to be similar to the stars'. Therefore, we
515: set $g_M(x) = C_M \exp(x)$,
516: yielding a BH number density of $C_M n_*$ near $r_i$. This is in
517: contrast to the work of BW77 and HA06a, who both assume that the
518: population of stars was already in thermal equipartition (i.e. $M
519: \beta_M = M_* \beta_*$). BW77 addressed the distribution of stars in
520: globular clusters, which have half-mass relaxation times much shorter
521: than the Hubble time and so the entire system should reach
522: equipartition. In contrast, in galactic nuclei the relaxation
523: timescale for large radii is usually longer than the age of the
524: universe, and so the source population could not reach a complete
525: equilibrium. HA06 made a numerical error in how they normalised the BH
526: distribution and so had reasonable agreement between their work and
527: previous works. However, with the proper normalisation, they would
528: have found the number of BHs to be $(M / M_*)^{3/2} = 10^{3/2}$ times
529: larger than their results indicated. This is because in the case of a
530: velocity dispersion independent of mass, the ratio of number densities
531: of objects is described by $n_M/n_* = C_M/C_*$, whereas for a
532: population in thermal equilibrium we get
533: $n_M/n_* = C_M M^{3/2}/(C_* M_*^{3/2})$.
534: Because the steady-state distribution function is most sensitive to
535: $g(0)$ and not the functional form $g(x<0)$, their results are similar
536: to ours when one uses the incorrect normalisation.
537:
538: One goal of our analysis is to see if the mass distribution of BHs may be
539: determined either through dynamical interactions with their environment, or
540: through the release of GWs. We therefore look at a variety of
541: time-independent BH mass functions parametrised as power-laws. The models
542: are determined by four basic parameters: the total number fraction of BHs,
543: $C_{\rm BH} = \sum_M C_M$; the slope of the BH mass function, $\beta$,
544: where $\rmd n_M / \rmd M \propto M^{-\beta}$; the minimum BH mass, $M_{\rm
545: min}$; and, finally, the maximum BH mass in the nucleus, $M_{\max}$. In
546: our calculations, we approximate the distribution of the BHs with $N=9$
547: discrete distribution functions with
548: \begin{equation}
549: \label{eq:bhdist}
550: C_i = \int_{M_{\rm min}+i \Delta M}^{M_{\rm min}+(i+1) \Delta M}
551: \gamma M^{-\beta} \rmd M,
552: \end{equation}
553: where $\Delta M = (M_{\max} - M_{\rm min})/N$, and $\gamma$ is
554: normalised such that $\sum_i C_i = C_{\rm BH}$. We set the mass for
555: each distribution function to be the average mass for that bin,
556: \begin{equation}
557: \label{eq:bhmass}
558: M_i = \int_{M_{\rm min}+i \Delta M}^{M_{\rm min}+(i+1) \Delta M}
559: (\gamma / C_i) M^{-\beta+1} \rmd M.
560: \end{equation}
561: We describe the numbers used above in detail below, and outline them in
562: Table~\ref{table1}.
563:
564: \begin{table*}
565: \centering
566: \begin{minipage}{140mm}
567: \caption{\label{table1} BH Models and Results. {\new Unless otherwise
568: noted, these results apply to a Milky Way like galaxy with
569: $\msmbh = 3.5\times 10^6\,\msun$, $\sigma_* = 75\,\kms$ and $r_i
570: = G \msmbh/\sigma_*^2 = 2.7\,$pc.} The columns are, from left to
571: right, (1) the model name, (2) the slope of the BH mass
572: function, (3) the minimum BH mass, (4) the maximum BH mass, (5)
573: the fraction of stars that are BHs, (6) the merger rate of GW
574: capture binaries per model galaxy, and (7) the expected AdLIGO
575: detection rate as calculated in \S~\ref{sec:detection}.}
576: \begin{tabular}{@{}lcccccc@{}}
577: \hline
578: Model & $\beta$ & $M_{\rm min}$ & $M_{\max}$ & $n_{\rm BH}/n_*$ & Merger Rate per Galaxy & AdLIGO Detection Rate \\
579: & & ($\Msun$) & ($\Msun$) & & (yr$^{-1}$) & ($\xi_{30}\yr^{-1}$) \\
580: \hline
581: A & 2 & 5 & 10 & 0.001 & $2.2\times 10^{-10} $ & 5.8 \\
582: A$\beta3$ & 3 & 5 & 10 & 0.001 & $2.0\times 10^{-10} $ & 5.2 \\
583: B & 2 & 5 & 15 & 0.001 & $2.8\times 10^{-10} $ & 16 \\
584: B$\beta3$ & 3 & 5 & 15 & 0.001 & $2.5\times 10^{-10} $ & 13 \\
585: B-1 & 2 & 5 & 15 & 0.1 & $5.3\times 10^{-9} $ & 220 \\
586: B-2 & 2 & 5 & 15 & 0.01 & $8.3\times 10^{-10} $ & 40 \\
587: Be5\footnote{This is evaluated for a SMBH with $\msmbh = 1\times 10^5\,\Msun$ and $\sigma_* =30\,\kms$.} & 2 & 5 & 15 & 0.001 & $2.8\times 10^{-10} $ & 15 \\
588: C & 2 & 5 & 25 & 0.001 & $3.8\times 10^{-10} $ & 46 \\
589: D & 2 & 10 & 15 & 0.001 & $3.4\times 10^{-10} $ & 23 \\
590: E & 2 & 10 & 25 & 0.001 & $3.2\times 10^{-10} $ & 12 \\
591: E-1 & 2 & 10 & 25 & 0.1 & $1.2\times 10^{-8} $ & 1200 \\
592: E-2 & 2 & 10 & 25 & 0.01 & $1.3\times 10^{-9} $ & 150 \\
593: F-1 & 2 & 10 & 45 & 0.1 & $1.5\times 10^{-8} $ & 2700 \\
594: BSR\footnote{There is an additional population of $25\,\Msun$ BHs with $n_{\rm bh}/n_* = 10^{-5}$} & 0 & 5 & 10 & 0.001 & $2.9\times 10^{-10} $ & 43 \\
595: \hline
596: \end{tabular}
597: \end{minipage}
598: \end{table*}
599:
600:
601: The number fraction of BHs in galactic nuclei is sensitive to the
602: initial mass function (IMF) of high mass stars. For a Kroupa IMF
603: {\new \citep{2003ApJ...598.1076K}}
604: we estimate $C_{\rm BH} \approx 0.001$ by assuming that all stars with mass
605: $M>20\,\Msun$ become BHs. A similar distribution of stars formed
606: uniformly throughout time was found to be consistent with the K-band
607: luminosity distribution of stars in the galactic centre
608: \citep{1999ApJ...520..137A}. However, more recent observations of our
609: Galactic Center suggest the IMF slope of high mass stars may be
610: considerably shallower than those in regular clusters \citep[see,
611: e.g.,][]{2005MNRAS.364L..23N, 2006MNRAS.366.1410N,2006ApJ...643.1011P,
612: 2007ApJ...669.1024M}. \citet{2006ApJ...643.1011P} found that the
613: stars in an apparent disk around Sgr A* are best fit with a flattened
614: IMF $\propto m^{-\beta_{\rm imf}}$ with $\beta_{\rm imf} \approx
615: 0.85-1.35$ and a depletion of low mass stars. This is consistent with
616: the work of \citet{2005MNRAS.364L..23N}, who constrained the number of
617: young low mass objects with X-ray observations, and determined there
618: to be a deficiency of stars with mass $\lesssim 3\,\Msun$. Most
619: recently, \citet{2007ApJ...669.1024M} looked at late-type giants and
620: concluded that the MF of high mass stars may be as shallow as
621: $\beta_{\rm imf} \approx 0.85$. A shallow MF is also expected
622: theoretically \citep{\morris} and is consistent with the most recent
623: hydrodynamic simulations of star formation in the
624: Galactic Center
625: \citep{2008ApJ...674..927A}. Considering the abundant evidence for an
626: alternative IMF in galactic nuclei, we therefore look at a wide range
627: of number fractions, setting $C_{\rm BH} = 0.001, 0.01,$ and $0.1$,
628: which roughly corresponds to $\beta_{\rm imf} = 2.3, 1.5,$ and $0.8$
629: respectively.
630:
631:
632: The mass distribution of BHs is relatively unconstrained. Currently,
633: the best constraints come from the couple of tens of X-ray binaries
634: with dynamically determined BH masses \citep{2006ARA&A..44...49R}.
635: Each measurement often has considerable uncertainty, but the masses
636: seem to span a range of $\sim 3 - 18\,\Msun$. There is now strong
637: dynamical evidence for BHs with even greater masses, up to
638: $23$--$34\,\Msun$,
639: in low metallicity environments
640: \citep{2007Natur.449..872O,2007ApJ...669L..21P,2008ApJ...678L..17S}. Nevertheless,
641: these observations suffer from a severe observational bias; the BHs
642: must be in a close binary to be observed. Theoretical estimates for
643: the mass distribution are also highly uncertain.
644: \citet{2004ApJ...611.1068B} have used their sophisticated population
645: synthesis models to determine the expected distribution of BH masses
646: in a variety of environments. Typically, they expect most BHs in high
647: metallicity environments to have a uniform distribution of masses
648: between $\sim 5 - 10\,\Msun$, but do find that significantly more
649: massive BHs may form from the merger of the BH with a high mass
650: companion star. Unfortunately, these massive BHs would not be found in
651: X-ray binaries, unless they were introduced into one dynamically.
652: Given these uncertainties, we choose to parametrise the mass
653: distribution of BHs as a power-law $dn_M/dm \propto m^{-\beta}$, where
654: we consider $\beta= 2$ and $\beta = 3$. We look at the importance of
655: the highest mass BHs, by modelling different upper limits on BH mass,
656: $M_{\max}$. We also have a model based on the work of
657: \citet{2004ApJ...611.1068B}. In Model BSR, we use a flat $\beta = 0$
658: model of BHs with masses between $5$ and $10\,\Msun$ with a fraction
659: $\sim 0.01$ of BHs with mass $25\,\Msun$. This is consistent with
660: their Model C2 of solar metallicity stars and based on their Figure~7.
661:
662:
663:
664:
665:
666:
667: For the remainder of \S~\ref{sec:massseg}, we focus our results on
668: Milky Way like nuclei, and assume that $\msmbh = 3.5\times
669: 10^6\,\Msun$ \citep{2005ApJ...620..744G,2005ApJ...628..246E} and that
670: the velocity dispersion of the stars and stellar BHs is $\sigma_* =
671: 75\,\kms$ (HA06). In \S~\ref{sec:GB} we also
672: consider other galaxies that have relaxation times short enough to
673: reach steady state in a Hubble time with $\msmbh \sim
674: 10^4-10^7\,\Msun$, in order to calculate the overall rate of GWs
675: sources in the universe.
676:
677:
678:
679:
680: \subsection{Results and Implications}
681: \label{sec:massseg:results}
682:
683: \begin{figure*}
684: \centering
685: \includegraphics[width=\columnwidth]{numberdenB.ps}
686: \includegraphics[width=\columnwidth]{numberdenE-2.ps}
687: \caption{\label{fig:numberdensity} The number density of stars and
688: BHs for Model B (left) and E-2 (right). The top dash--dotted line is the
689: number density of the stars as a function of radius. The
690: alternating dotted and dash-dotted lines show the number density
691: of the separate mass bins used in our calculations with the most
692: massive BHs having the largest number density interior to $r
693: \approx 0.3-0.6\,$pc. The solid blue line is the total BH number
694: density $=\sum_M n_M(r)$. Near $r \approx 1\,$pc the potential of
695: the stars and BHs is equal to the Keplerian potential of the SMBH.
696: The rapid drop in the number of BHs drops rapidly here because of
697: the rigid boundary condition at $\phi = 0$. In all of our
698: simulations, the bin of the most massive BHs dominates the number
699: density of BHs in the inner $\sim 0.1\,$pc. }
700: \end{figure*}
701:
702: For all of our models, independent of the BH mass function (namely,
703: $\beta$, and the mass range of BHs), the most massive BHs always become the
704: dominant BH species in the inner $\sim 0.1\,$pc of the galactic nucleus. In
705: Figure~\ref{fig:numberdensity}, we plot the number density of the stars and
706: BHs as a function of radius. As has been found previously (BW77, HA06,
707: FAK06), deep in the Keplerian potential of the SMBH, the distribution
708: functions of the stars and BHs become power-laws of the negative specific
709: energy of the objects $g_m(x) \propto x^{p_m}$, and hence have a power-law
710: density profile $\propto r^{-3/2-p_m}$. We find that throughout all of our
711: simulations, the exponent $p_m$ is best fit by a linear relationship
712: between the mass ratio of the object, $m$, and the most massive BH, $M_{\rm
713: max}$,
714: \begin{equation}
715: \label{eq:powerlawrelation}
716: p_m = p_0 \frac{m}{M_{\max}},
717: \end{equation}
718: where $p_0 \approx 0.5-0.6$. We have found that $p_0$ usually has a
719: small scatter ($\sim 20\%$), depending on $M_{\max}$ and $\beta$;
720: however, given $p_0$, Eq.~(\ref{eq:powerlawrelation}) is often
721: accurate to $\lesssim 1\%$. This relationship is similar to that found
722: in BW77, who found that $p_0 \approx 0.25-0.3$, when they looked at
723: two different components with comparable number densities.
724: \citet{2008arXiv0808.3150A} attributes the steeper density profiles to
725: ``strong'' mass segregation, where the relaxation of the system is
726: determined by the many low mass objects. In their calculations, $p_0$
727: increases monotonically with $M_{\max}$, however the rate at which
728: it increases is rather slow after $p_0 \sim 0.5$. In contrast, in our
729: simulations with a mass spectrum of massive objects and accounting for
730: the loss-cone, we do not find that the maximum power-law index $p_0$
731: to depend sensitively on the number or mass of the highest mass
732: objects. However, we have not looked at the large range of parameters
733: of \citet{2008arXiv0808.3150A}.
734:
735:
736: Mass segregation in galactic nuclei ceases when the BHs begin to dominate
737: the relaxation process in the inner cusp. Typically, this is expected when
738: $n_{\rm BH} M_{\rm BH}^2 > n_* M_*^2$. One therefore may try to detect the
739: presence of a cluster of BHs through their interactions with the luminous
740: stars in its vicinity. We can estimate the relaxation timescale at a
741: radius $r$ from Eqs.~(\ref{eq:relaxtime})~\&~(\ref{eq:taur}), in terms of
742: the Keplerian potential of the SMBH
743: \begin{equation}
744: \label{eq:relaxtimecusp}
745: t_r(r) = \frac{3 (2 \pi v_c^2(r))^{3/2}}{32 \pi^2 G^2 \ln{\Lambda}
746: \sum_M{M_M^2 n_M(r)}}
747: \end{equation}
748: where $v_c(r) = \sqrt{G \msmbh / r}$ is the circular velocity around
749: the SMBH at radius $r$. In Figure~\ref{fig:relax}, we have plotted
750: the relaxation timescale as a function of radius for Models B and E-2.
751: The BHs begin to dominate the local relaxation processes at $r \sim
752: 0.5\,$pc, consistent with the results of FAK06 and HA06. We find that
753: the relaxation timescale in the galactic centre is highly sensitive to
754: the assumed mass distribution of the BHs, since the lower density of
755: the BHs is easily outweighed by their greater mass. By finding a
756: tracer of the relaxation time in the inner $0.1\,$pc of the galactic
757: centre, it may be possible to determine if such heavy BHs even exist
758: in our galaxy \citep{2002ApJ...571..320C}.
759:
760: \begin{figure*}
761: \centering
762: \includegraphics[width=\columnwidth]{relaxB.ps}
763: \includegraphics[width=\columnwidth]{relaxE-2.ps}
764: \caption{\label{fig:relax} The relaxation timescale as a function of
765: radius for Model B (left) and E-2 (right) plotted in Gyr. The
766: solid line shows the relaxation timescale for the entire system.
767: The long dashed and dotted lines show the contribution of
768: the BHs and stars respectively. Although the stars contribute to
769: the formation of the cusp and are the most common objects in the
770: system, their low mass precludes them from dominating the
771: relaxation in the inner $\approx 0.6\,$pc of the cusp. The break
772: in the relaxation time at $\sim 0.0002\,$pc is due to our boundary
773: condition that no stars have $x > 10^4$. }
774: \end{figure*}
775:
776: As discussed above, a population of massive BHs in the galactic centre
777: may be revealed through strong encounters with their neighbouring
778: stars. Because the BHs are more massive than the typical star, they
779: tend to drive stars out of the nucleus, creating a shallow stellar density
780: profile (FAK06; HA06; for observations see
781: \citealt{2003ApJ...594..812G} and more recently
782: \citealt{2007AA...469..125S}). \citet{\meg} proposed that a
783: population of such BHs may be seen through stars on very radial orbits
784: that are ejected by close encounters by the BHs. In some extreme
785: encounters, \citet{2008MNRAS.383...86O} showed that strong
786: interactions between BHs and stars can lead to the stars being ejected
787: from the Milky Way altogether as the observed hypervelocity stars
788: \citep{2005ApJ...622L..33B,2006ApJ...640L..35B}. The results of
789: \citet{2008MNRAS.383...86O} were sensitive to the assumed BH mass
790: distribution. Self-segregation of the BHs is an important
791: consideration when evaluating the likelihood of the scenario, and the
792: velocities of ejected HVSs \citep{2007MNRAS.379L..45S}.
793:
794: As a more extreme example of a strong encounter, BHs and stars may
795: physically collide into each other \citep[][FAK06]{\morris} producing
796: strong X-ray and UV flares in galactic nuclei. Interestingly, we find
797: that the total rate of physical collisions between stars and BHs can
798: be comparable to the rate stars are tidally disrupted by the
799: SMBH, and may be visible in variability surveys of galactic nuclei.
800: Such encounters may have smaller bolometric luminosities than stars
801: disrupted by the SMBH due to their lower Eddington luminosity.
802:
803:
804: \section{Binary Formation, Inspiral, and Event Rates}
805: \label{sec:GB}
806:
807: When two compact objects have a close encounter, they can emit
808: sufficient energy through GWs that they become bound and form a
809: binary. {\new To see the importance of such a process in galactic
810: nuclei, we first estimate the rate of binary formation for two
811: typical $10\,\msun$ BHs. We present more detailed calculations
812: later in this section. To first order, two BHs with an initial
813: relative velocity $w$ will form a
814: bound binary due to the emission of GWs
815: if they come within a distance
816: $\approx 7.4\times 10^{-3}\,\rsun (w/100\,\kms)^{-4/7}$ at closest
817: approach (see Eq.~[\ref{eq:rpmax}] below). This corresponds to an initial impact parameter (the
818: distance of closest approach if the BHs moved in straight lines) of (Eq.~[\ref{eq:bmax}])
819: $b \la 2.4\,\rsun (w/100\,\kms)^{-9/7}$. We expect that a single BH
820: will undergo such an encounter approximately every $(\pi b^2 n_{\rm
821: BH} w)^{-1} = 3.4\times 10^{12} (n_{\rm BH}/10^6 {\rm
822: pc}^{-3})^{-1} (w/100\,\kms)^{11/7}\,\yr$. The total rate of
823: binary formation is then approximately (see, e.g., Eq.~[\ref{eq:diffsimpfin}]) $\int n_{\rm BH}^2 \pi
824: b^2 w 4 \pi r^2 \rmd r$, where $w(r) \sim v_c(r) = \sqrt{G \msmbh/r}$
825: is the Keplerian circular velocity at radius $r$. Integrating from
826: $r = 10^{-3}\,$pc to $r = 1 \,$pc, yields a rate of $3.4\times
827: 10^{-10}\,\yr^{-1} [N_{\rm BH}/ (2\times 10^4)]^2$ for $N_{\rm BH}$ BHs in
828: a density cusp around the SMBH with $n_{\rm BH} \propto
829: r^{-2}$. This is in good agreement with our more detailed
830: calculations presented in \S~\ref{sec:rates}.
831:
832: In this section, we calculate the formation rate of such
833: gravitational wave capture binaries in greater detail. {\new We
834: first calculate the conditions necessary to form gravitational wave
835: capture binaries and the properties of their eventual inspiral in
836: different environments. We then calculate the rate of binary
837: formation for the multi-mass, segregated clusters of BHs that we
838: modelled in \S~\ref{sec:massseg}. Finally, we extrapolate our
839: results for galactic models, including the distribution of SMBHs
840: in the universe and accounting for the statistical variance in the
841: number density of stars surrounding the SMBHs, in order to estimate the
842: comoving rate density of binary formation and inspiral. }
843:
844: In our calculations, we express our equations with the symmetric mass
845: ratio of the encounter $\eta = (m M)/ (m+M)^2$, and total mass of the
846: BHs $\mtot$. With these variables, the reduced mass is $\mu = \mtot
847: \eta$.
848:
849: }
850:
851:
852: \subsection{Orbital Evolution}
853:
854: In the following sections, we shall work with dimensionless parameters
855: in units ${G} = {c} = 1$, where the total mass $\mtot$ has length and time
856: dimensions. We define the dimensionless pericenter distance as
857: $\rho_{p}=r_p/\mtot$, dimensionless semimajor axis for an eccentric orbit
858: with eccentricity $e$ as $\alpha\equiv a/\mtot = \rho_p/(1-e)$.
859: In these units, the mean orbital angular frequency from Kepler's law is
860: simply $\omega_{\rm orb}=\mtot^{-1} \alpha^{-3/2}=\rho_p^{-3/2}(
861: 1-e)^{3/2}$, the angular frequency at pericenter passage is
862: $\omega_{p}=\mtot^{-1} \rho_p^{-3/2}(1 +e)^{1/2}$, the orbital time is
863: $\Delta t_{\rm orb}\equiv f_{\rm orb}^{-1}\equiv 2\pi\omega_{\rm
864: orb}^{-1} $. We define the characteristic time of pericenter passage
865: as $\Delta t_{p}\equiv f_{\rm p}^{- 1}\equiv 2\pi\omega_{p}^{-1}$,
866: which satisfies $\Delta t_{p}=\Delta t_{\rm orb}$ for circular
867: encounters, but is much smaller for eccentric encounters.
868:
869: \subsubsection{First Passage, Formation of Binaries}
870:
871:
872: \label{sec:binform}
873: Two BHs of mass $m$ and $M$ will form a binary if they undergo a close
874: encounter and release enough energy to become bound, $\delta E > \eta \mtot w^2/2$, where
875: $w=|\bm{v}_m-\bm{v}_M|$ is the magnitude of relative velocity of the BHs at infinity,
876: and we ignore the presence of the SMBH during the encounter and assume that the
877: interaction is local. Given the relativistic nature of such events,
878: and the comparatively low velocity dispersions in nuclei, the
879: encounters are always nearly parabolic
880: \citep{1987ApJ...321..199Q,1993ApJ...418..147L}. In this limit, the
881: amount of energy released during the encounter is
882: \citep{1963PhRv..131..435P, 1977ApJ...216..610T}
883: \begin{eqnarray}
884: \label{eq:de}
885: \delta E
886: &\approx&
887: -\frac{85\pi}{12\sqrt{2}} \frac{\eta^2\mtot^{9/2}}{r_{p}^{7/2}},
888: \end{eqnarray}
889: where $r_{p}$ is the distance of closest approach,
890: \begin{eqnarray}
891: r_{p} &=& \left(\sqrt{\frac{1}{b^2}+\frac{\mtot^2}{b^4 w^4}} +
892: \frac{ \mtot}{b^2 w^2}\right)^{-1}\nonumber\\
893: &\approx& \frac{b^2 w^2}{2 \mtot}\left( 1 - \frac{b^2 w^4}{4\mtot^2}\right)
894: \label{eq:rmin}
895: \end{eqnarray}
896: and $b$ is the impact parameter of the encounter, and in the second line we have expanded to first order in $w/c$.
897:
898: The final properties of the system are determined by the system's final energy,
899: $E_{\rm final}= \mtot \eta w^2/2+ \delta E$, and angular momentum,
900: $L_{\rm final} = \mtot \eta b w + \delta L$, where
901: \begin{eqnarray}
902: \label{eq:deltal}
903: \delta L &\approx& -\frac{6\pi \mtot^4\eta^2}{r_p^2},
904: \end{eqnarray}
905: is the amount of angular momentum lost in GWs
906: \citep{1964PhRv..136.1224P}. For nearly all of the encounters we find
907: $|\delta L| \ll \mtot \eta b w $ and so we set $\delta L = 0$. If the total
908: energy of the system is negative, $E_{\rm final} < 0$, then the system
909: will remain bound with a semi-major axis,
910: \begin{equation}
911: \label{eq:a}
912: a_0 = - \frac{\mtot^2\eta}{2 E_{\rm final}},
913: \end{equation}
914: eccentricity,
915: \begin{equation}
916: \label{eq:e}
917: e_0 = \sqrt{1 + 2 \frac{E_{\rm final} b^2 w^2}{\mtot^3 \eta}},
918: \end{equation}
919: and pericenter, $r_{p0} = a_0 (1-e_0)$.
920:
921: With the criterion $E_{\rm final} < 0$ for binary formation, the
922: maximum impact parameter for two BHs with relative velocity $w$ to
923: become bound is
924: \begin{eqnarray}
925: \label{eq:bmax}
926: b_{\max} &=& \left(\frac{340\pi}{3}\right)^{1/7} \mtot \frac{\eta^{1/7}}{w^{9/7}}.
927: \end{eqnarray}
928: Substituting in Eq.~(\ref{eq:rmin}), this corresponds to a maximum pericenter distance
929: \begin{eqnarray}
930: \label{eq:rpmax}
931: r_{p,\max} &=& \left( \frac{85 \pi}{6\sqrt{2}}\right)^{2/7} \mtot
932: \frac{\eta^{2/7}}{w^{4/7}}\\\nonumber
933: &&\quad\times
934: \left(1-\frac{1}{4}\left( \frac{85 \pi}{3}\right)^{2/7}(4\eta)^{2/7}w^{10/7}\right).
935: \end{eqnarray}
936: We shall demonstrate below that event rates are dominated by the central regions of the
937: galactic cusp.
938: For $m = M$ ($\eta =1/4$) and $w= 1000\,\kms$, $b_{\max} \approx 2900\,\mtot$, and $r_{p{\max}}= 47\,\mtot$.
939: The correction in Eq.~(\ref{eq:rpmax}) is clearly negligible for nonrelativistic initial conditions.
940:
941: \subsubsection{Eccentric Inspiral}
942: After the binary forms with dimensionless pericenter distance $\rho_{p0}=r_{p0}/\mtot$ and eccentricity $e_0$, its orbit will decay through the emission of
943: GWs \citep{\peters},
944: \begin{equation}
945: \frac{\rmd \rho_p}{\rmd t} =-\frac{64}{5} \frac{\mtot^{-1} \eta^2}{\rho_p^{3}}\frac{(1-e)^{1/2}}{(1+e)^{7/2}}\left(1+\frac{73}{24} e^{2}+\frac{37}{96} e^{4}\right)
946: \label{pet1}
947: \end{equation}
948: \begin{equation}
949: \frac{de}{dt} = -\frac{304}{15} \frac{\mtot^{-1} \eta}{\rho_p^{4}}e\frac{(1-e)^{3/2}}{(1+e)^{5/2}}\left(1+\frac{121}{304}e^{2}\right)
950: \label{pet2}
951: \end{equation}
952: where we assume $\rho_{p}=r_p/\mtot= a/[\mtot(1-e)]$ until the last stable orbit.
953:
954: We evolve the orbits following \citet{\peters} for the
955: evolution of the dimensionless periapsis $\rho_p$, the
956: time-to-merger, $t$, as a function of the instantaneous eccentricity,
957: $e$. Note that if starting from a parabolic passage, the initial condition
958: for the orbit is determined from exactly one parameter, $\rho_{p0}$.
959: Dividing Eq.~(\ref{pet1}) with Eq.~(\ref{pet2}) yields a separable differential
960: equation for $\rho_p(e)$. The solution satisfying a parabolic encounter initial
961: condition $\rho_{p}=\rho_{p0}$ at $e=1$ is
962: \begin{equation}\label{e:rho_p1}
963: \rho_p(\rho_{p0},e) = \rho_{p0} \kappa_{\rho}(e)
964: \end{equation}
965: where
966: \begin{equation}\label{e:rho_p2}
967: \kappa_{\rho}(e) = 2 \left(\frac{304}{425}\right)^{\frac{870}{2299}} e^{\frac{12}{19}}\left(1+\frac{121}{304}e^2\right)^{\frac{870}{2299}}
968: (1+e)^{-1}
969: \end{equation}
970: for which $\kappa_{\rho}=1$ at $e=1$. The orbital evolution
971: equations~(\ref{e:rho_p1}) \& (\ref{e:rho_p2}) are valid far from the
972: horizon $\rho_p\gg 2$. Once the last stable orbit (LSO) is reached the
973: evolution is no longer quasiperiodic and the binary quickly coalesces. In the
974: leading order ratio approximation for {\new infinite mass ratio and} zero
975: spins,
976: \begin{equation}
977: \label{eq:LSO}
978: \rho_p(e_{\rm LSO}) = \frac{6+2e_{\rm LSO}}{1+e_{\rm LSO}},
979: \end{equation}
980: which we solve numerically for $e_{\rm LSO}(\rho_{p0})$ using
981: Eqs.~(\ref{e:rho_p1}) \& (\ref{e:rho_p2}).
982:
983: The time evolution of the eccentricity can be written using Eqs.~(\ref{pet1})~\&~(\ref{pet2}) as
984: \begin{eqnarray}
985: t_{\rm mg}(e) &=& \frac{15}{19} \left(\frac{304}{425}\right)^{\frac{3480}{2299}} \mtot \eta^{-1} \rho_{p0}^4 \nonumber\\
986: &&\quad\times \int_0^{e} {\epsilon}^{\frac{29}{19}} \frac{\left(1+\frac{121}{304}{\epsilon}^2\right)^{\frac{1181}{2299}}}{\left(1-{\epsilon}^2\right)^{3/2}} \D \epsilon
987: \label{e:t-e}
988: \end{eqnarray}
989: where $t_{\rm mg}(e)$ denotes the time remaining until coalescence when the eccentricity is $e$.
990: Close to coalescence this is approximately,
991: \begin{equation}
992: t_{\rm mg}(e) \approx \frac{5}{16} \left(\frac{304}{425}\right)^{\frac{3480}{2299}} \mtot \eta^{-1} \rho_{p0}^4 e^{\frac{48}{19}} {~~\rm if}~ e\ll 1.
993: \end{equation}
994: while the total merger time starting from a highly eccentric initial orbit \citep{\peters}
995: \begin{equation}
996: t_{\rm mg}\approx \frac{3}{85} \frac{a^4}{\mtot^3 \eta} (1-e^2)^{7/2} \text{~~if~} e\approx 1.
997: \label{timerge}
998: \end{equation}
999:
1000: Substituting Eqs. (\ref{eq:a}) \& (\ref{eq:e}) for the initial separation and eccentricity $a_0$ and $e_0$
1001: into Eq. (\ref{timerge}) we get,
1002: \begin{equation}
1003: \label{eq:fimerge}
1004: t_{\rm mg} \approx \frac{3 \sqrt{3}}{170 \sqrt {85 \pi}}
1005: \frac{(b w )^{21/2} }{\mtot^{19/2} \eta^{3/2} }
1006: \end{equation}
1007: where we have set $E_{\rm final} = \delta E$, ignoring the initial energy
1008: of the binary. This is an excellent approximation all the way out to $0.99
1009: b_{\rm bmax}$ because of the strong dependence of $t_{\rm merge}$ on the
1010: initial velocity $w$.
1011: The merger time reaches its maximum at $b = b_{\max}$ and simplifies to
1012: $4\pi G \mtot / w^3$: about two times the orbital period of a binary with
1013: circular velocity $w$.
1014:
1015: For the most relevant systems, we find that such binaries will merge before
1016: they have a close encounter with a third body. The binary can be disrupted
1017: if it encounters a third body with a closest approach $\lesssim 2a$ before
1018: it merges. For the maximum impact parameter to form a binary, $b_{\rm
1019: max}$, the separation of the binary that forms is $a \approx G \mtot /w^2$.
1020: The typical timescale for an encounter to disrupt the binary is $\sim (12
1021: \pi a^2 n w)^{-1} = w^3/(12 \pi G^2 \mtot^2 n)$, where the additional
1022: factor of $3$ comes from gravitational focusing. Because of the strong
1023: dependence Eq.~(\ref{eq:fimerge}) has on the impact parameter ($\propto
1024: b^{21/2}$), for any reasonable disruption timescale, a slightly closer
1025: impact will always compensate and allow the binary to merge before being
1026: disrupted. In fact, in order for the disruption timescale to be comparable
1027: to merger timescale, the number density of stars must be $n \gtrsim
1028: 10^{15}\,$pc$^{-3}$ for $w = 100\,\kms$. Because every binary results in a
1029: rapid merger, the rate of inspiral at any time is well approximated by the
1030: rate of binary formation.
1031:
1032: Equation~(\ref{e:t-e}) provides an explicit monotonic relation between
1033: time and eccentricity. This relation is sensitive to the free
1034: parameters $\mtot$, $\eta$, and $\rho_{p0}$ only through the
1035: overall scaling $\mtot \eta^{-1} \rho_{p0}^4$. Therefore
1036: eccentricity can be thought of as a dimensionless time
1037: variable. Figure~\ref{f:t-e} illustrates the conversion between $t$
1038: and $e$. For the parameters, $\rho_{p0}=40$, $\mtot=20\Msun$, and
1039: $\eta=0.25$, the characteristic timescale is several hours for
1040: moderate--large eccentricities, and minutes for the final small
1041: eccentricity inspiral.
1042: \begin{figure}
1043: \centering
1044: \mbox{\includegraphics{time_eccentricity.eps}}
1045: \caption{\label{f:t-e} The eccentricity evolution as a function of time to
1046: merger. All of the orbits start off near $e=1$, and depending on the
1047: initial periapsis $\rho_{p0}$, LSO corresponds to the eccentricity $e_{\rm
1048: LSO}$ shown with filled circles as marked. Open circles give $e_{\rm LSO}$
1049: for equidistant intermediate values for $\rho_{p0}$. The curves should not
1050: be extrapolated to $e< e_{\rm LSO}$ as the evolution might be very
1051: different during the BH merger.}
1052: \end{figure}
1053:
1054:
1055:
1056: \subsection{Event Rates}
1057: \label{sec:rates}
1058: Given the cross-section for binary formation ($\sigma_{\rm cs}=\pi
1059: b_{\max}^2$), we can now calculate the expected rate of binary
1060: formation in a single galactic nucleus, $\Gamma_{1\rm GN}$. In order
1061: to evaluate the total contribution of different BH masses, $(m,M)$,
1062: velocities, $(\bm{v}_m,\bm{v}_M)$, and spatial position, $r$, within a
1063: single galactic nucleus, we need to integrate the differential rate
1064: $\D^9 \Gamma_{1\rm GN}=\sigma_{cs} w f_m(r,\bm{v}_m)f_M(r,\bm{v}_M)$
1065: over the corresponding distributions. Here $w = | \bm{v}_m - \bm{v}_M
1066: |$ is still the magnitude of relative velocity between the BHs, and
1067: $f_m$ and $f_M$ are the 6-dimensional distribution functions of the
1068: BHs with mass $m$ and $M$ derived in \S~\ref{sec:massseg}.
1069: Generally, this calculation requires us to evaluate an integral over 9
1070: variables,
1071: \begin{eqnarray}
1072: \label{eq:integral}
1073: \Gamma_{1\rm GN} &=& \int_{r_{\min}}^{r_{\max}}\rmd r 4\pi r^2 \int_{M_{\min}}^{M_{\max}} \rmd M
1074: \int_{M_{\min}}^{M} \rmd m \\\nonumber
1075: && \quad \int\!\!\!\int_{x_m, x_M > 10, J >J_{\rm LC}} \hspace{-1.8cm}\rmd^3 v_m \rmd^3 v_M f_m(r,\bm{v}_m)f_M(r,\bm{v}_M) \sigma_{\rm cs} w,
1076: \end{eqnarray}
1077: where the integration bounds are set consistent with
1078: the distribution domains (see \S \ref{sec:massseg}). The multi-dimensional
1079: integration can be greatly simplified to only three variables using a
1080: few approximations. First in \S 2.3, we assumed that the the velocity distributions
1081: were isotropic. For fixed $m$ and $M$, we switch integration variables from
1082: $(\bm{v}_m, \bm{v}_M)$ to $(\bm{v}_m-\bm{v}_M, \bm{v}_m+\bm{v}_M)$, and adopt
1083: spherical coordinates. Since the integrand only depends on $w$, we can evaluate the
1084: integrals over the other 5 velocity components\footnote{Some difficulties
1085: arise because of the integration bounds depend on the other parameters. However,
1086: the integrals can be evaluated under the approximation $r_{\min}\leq r \leq r_{\max}$ and $0\leq v\leq v_{\rm esc}(r)$,
1087: where $r_{\min}$ and $r_{\max}$ are the minimum and maximum radii for a relaxed population of BHs in the nucleus
1088: and $v_{\rm esc}(r)$ is the escape velocity at radius $r$.}
1089: \begin{eqnarray}
1090: \int \rmd^3 v_m \int\rmd^3 v_M\, f_m(r,\bm{v}_m)f_M(r,\bm{v}_M) \,\sigma_{\rm cs} w= \nonumber \\
1091: =n_m(r) n_M(r) \int \rmd w \, \psi_{mM}(r,w)\, \sigma_{\rm cs} w,
1092: \label{eq:relvel}
1093: \end{eqnarray}
1094: where $\psi_{mM}(r,w)$ is the relative velocity distribution which can be
1095: expressed as a one-dimensional integral using the Dirac-$\delta$ function for a power-law
1096: dimensionless energy distribution profile $g_m(x)\propto x^{p_m}$ (see \S~\ref{sec:massseg}
1097: for a definition). {\new Note} that the phase space distribution
1098: functions of the BHs and stars are well approximated by a power-law
1099: $g_m \propto x^{p_m}$ for $x > 10$.
1100:
1101: We have numerically integrated the remaining velocity integral
1102: over the possible range of $w$ for a variety of
1103: different slopes $p_m$ and $p_M$ and find that to within $\lesssim
1104: 10\%$ the integrand is independent of the shape of the {\new relative} velocity distribution and
1105: only depends on the expected value of the relative velocities, and can
1106: be expressed with the expected value of the individual velocity magnitudes as
1107: \begin{equation}
1108: \label{eq:simpint}
1109: \int \rmd w \, \psi_{mM}(r,w) \,\sigma_{\rm cs} w
1110: \approx \pi b_{\max}^2 v_{\rm c}(r),
1111: \end{equation}
1112: where $b_{\max}^2$ is evaluated at $w = v_{\rm c}(r)$ using Eq.~(\ref{eq:bmax}), and $n_m(r) \propto r^{-p_m-3/2}$.
1113: This is to be expected, since the encounters are practically parabolic such that the initial
1114: velocity profile is negligible compared to the velocity at periastron.
1115: Thus, Eq. (\ref{eq:integral}) simplifies to
1116: \begin{equation}
1117: \label{eq:diffsimpfin}
1118: \frac{\rmd^3 \Gamma_{1\rm GN}}{\rmd r \,\rmd m \,\rmd M}= 4 \pi^2 b_{\max}^2 v_{\rm c}(r) n_m(r)
1119: n_M(r) r^2,
1120: \end{equation}
1121: so that the total rate in one galactic nucleus is
1122: \begin{equation}
1123: \label{eq:intsimpfin}
1124: \Gamma_{1\rm GN} = \int_{r_{\min}}^{r_{\max}} \!\!\rmd r \int_{M_{\min}}^{M_{\max}} \!\!\rmd M \int_{M_{\min}}^{M} \!\!\rmd m
1125: \;\frac{\rmd^3 \Gamma_{1\rm GN}}{\rmd r \,\rmd m \,\rmd M}.
1126: \end{equation}
1127: In practice, we calculate $\Gamma$ for all of our simulations by calculating
1128: $n_m(r)$ and $n_M(r)$ for discreet masses and sum over $m < M$.
1129:
1130: \subsection{Results and Discussion}
1131: \label{sec:properties}
1132:
1133: The resulting event rates integrated over a single galactic nucleus
1134: are presented in the sixth column of Table~\ref{table1} for all of our
1135: model nuclei. The estimated rates vary between
1136: $\sim 10^{-8}$--$10^{-10}\yr^{-1}$ over the various models.
1137: We discuss the primary sources of uncertainties and other
1138: important aspects related to the event rates in detail below.
1139:
1140:
1141: \subsubsection{Number of Black Holes}
1142: Overall, we find that the merger rate is most sensitive
1143: to $C_{\rm BH} = n_{\rm BH}/ n_*$, however, the {\new accretion of BHs by
1144: the SMBH} reduces the na\"{i}ve scaling relation $\propto
1145: C_{\rm BH}^2$. Thus, as the number fraction of BHs increases by a
1146: factor of $100$, the rate of mergers only increases by a factor of
1147: $20$. The merger rate depends far less so on $M_{\max}$ and
1148: $M_{\rm min}$, since, for larger $M_{\max}$ the total number of BHs
1149: tends to be reduced in the inner $\sim 0.1\,$pc of the SMBH. The
1150: rate, however, remains relatively unchanged since the cross-section for
1151: binary capture increases with the BH mass.
1152:
1153: \subsubsection{Pericenter Distance Dependence}
1154: \begin{figure*}
1155: \begin{center}
1156: \includegraphics[width=\columnwidth]{dpdrp.ps}
1157: \includegraphics[width=\columnwidth]{gdpdrp.ps}
1158: \end{center}
1159: \caption{\label{fig:distribution} Normalised probability distribution
1160: and integrated probability distribution of initial pericenter
1161: distances for binaries in galactic nuclei (a) and {\new in
1162: Spitzer-unstable} massive star clusters {\new without a central
1163: massive black hole} (b{\new ; see \S~\ref{sec:clusters} for
1164: details}). To calculate the values in this figure, we have
1165: assumed that the merger rate is dominated by the $10\,\Msun$ BHs
1166: with a near Maxwellian velocity dispersion, and in galactic nuclei
1167: they have density profiles $\propto r^{-2}$, consistent with the
1168: density profile of the most massive objects in the nucleus (see
1169: \S~\ref{sec:massseg:results}). The solid line is the differential
1170: rate of binary formation $\rmd \Gamma / \rmd r_p$ and the dotted
1171: line represents the integrated probability $\Gamma (r>r_p)$. For
1172: galactic nuclei, the probability distribution is perfectly uniform
1173: out to pericenter distance $r_{p\max}(w_{\max}(.001\,$pc$) \approx
1174: 12 \mtot$, after which it drops off quickly. For globular clusters
1175: {\new without a central massive BH} the distribution is uniform out
1176: to $r_{p}\approx 250 \mtot$, and it begins to drop off as a Gaussian
1177: profile for $r_p \ga 540 \mtot$. }
1178: \end{figure*}
1179:
1180: For parabolic encounters, the differential of pericenter distances that
1181: lead to binary formation, according to Eq.~(\ref{eq:rmin}) is to leading order
1182: \begin{equation}
1183: \rmd r_p \approx \frac{ w^2 b \rmd b}{\mtot},
1184: \end{equation}
1185: where we have fixed $w$, the relative velocity at infinity. Since
1186: $\rmd \Gamma / \rmd b \propto b \propto \rmd r_p/ \rmd b $, the
1187: pericenter distribution {\new that lead to binary capture ($\rmd \Gamma / \rmd
1188: r_p$) } is uniform out to a maximum pericenter distance $r_{p{\max}}$. We can calculate the overall distribution of encounters
1189: that form binaries by changing the order of integration of
1190: Eq.~(\ref{eq:integral}) and leaving it as a function of $r_p$,
1191: \begin{eqnarray}
1192: \frac{\rmd \Gamma_{1\rm GN}}{\rmd r_p} &=& \int \!\!\! \int \!\!\! \int
1193: \!\!\! \int \rmd r \rmd w\rmd m \rmd M\; 4 \pi r^2 n_M(r) n_m(r) \nonumber\\&&\quad\times\frac{\mtot}{w^2} w \psi_{mM}(r,w),
1194: \label{eq:dgdrp}
1195: \end{eqnarray}
1196: where $\psi_{mM}(r,w)$ is the distribution function of relative velocities given by Eq.~(\ref{eq:relvel}).
1197: The limits of the integration determine the functional form in
1198: three main regimes. For $r_p < r_{p{\max}}[w_{\max}(r_{\min})]$ where
1199: $w_{\max}(r)= 2 v_{\rm esc}(r) = 2\sqrt{2} v_c(r)$
1200: is the maximum relative velocity at radius $r$, $v_{\rm esc}$ is the escape velocity, and $r_{p{\max}}(w)$ is defined in Eq.~(\ref{eq:rpmax}),
1201: Eq.~(\ref{eq:dgdrp}) is independent of $r_p$ and is integrated over
1202: the bounds
1203: \begin{align}
1204: 0 < w {} < w_{\max}(r) \nonumber \\
1205: r_{\rm min} < r {} < r_{\max}.
1206: \end{align}
1207: For $r_{p\max}(w_{\max}(r_{\rm min})) <r_p< r_{p\max}(w_{\max}(r_{\max}))$
1208: the limits of integration are determined by $w_{\rm
1209: eq}(r_p)$, the inverse of Eq.~(\ref{eq:rpmax}) and $r_{\rm eq}(r_p) = 2 G
1210: \msmbh/ w_{\rm eq}(r_p)^2$. In this regime, the limits of integration are
1211: split into two regions
1212: \begin{align}
1213: 0 < w < w_{\rm eq}(r_p) \nonumber \\
1214: r_{\rm min} < r {} < r_{\rm eq}(r_p),
1215: \end{align}
1216: and
1217: \begin{align}
1218: w_{\rm eq}(r_p) < w {} < w_{\max}(r) \nonumber \\
1219: r_{\rm eq}(r_p) < r {} < r_{\max}.
1220: \end{align}
1221: Finally, for $r_p > r_{p{\max}} (2 \sqrt{2} v_c(r_{\max}))$ the
1222: limits of integration are
1223: \begin{align}
1224: 0 < w {} < w_{\rm eq}(r_p) \nonumber \\
1225: r_{\rm min} < r {} < r_{\max}.
1226: \end{align}
1227: A normalised probability distribution of pericenter distances for
1228: encounters in galactic nuclei is plotted in
1229: Figure~\ref{fig:distribution}a, where we approximated $\psi_{mM}(r,w)$ as a
1230: Maxwellian distribution with variance $v_c(r)$.
1231: For comparison the distribution function for the Maxwellian core of a
1232: star cluster {\new without a central massive BH} is plotted in Figure~\ref{fig:distribution}b.
1233:
1234:
1235:
1236: \subsubsection{Eccentricity dependence}
1237:
1238:
1239: We can use Eqs.~(\ref{e:rho_p1}), (\ref{e:rho_p2}), and (\ref{e:t-e})
1240: to follow the secular evolution of the binary as it passes through the
1241: LIGO band and merges when it reaches its last stable orbit. In
1242: Figure~\ref{fig:inspirala}, we have plotted the secular evolution of
1243: formed binaries in $10\%$ probability intervals, and plotted the
1244: eccentricity after approximately every complete orbit with small
1245: circles. Nearly $90\%$ of all binaries actually form within the LIGO
1246: band, and $\approx 94\%$ have an eccentricity $e>0.3$ as it enters the
1247: LIGO band (when $f_p=10\Hz$). This signature is unique to an active
1248: cluster of BHs in a high velocity dispersion environment. For lower
1249: velocity dispersions as in star clusters, the {\new merger}
1250: distribution of pericenter distances {\new ($\rmd \Gamma/ \rmd r_p$)}
1251: stays uniform out to $\approx r_{p{\max}}(2\times 50\,\kms) \approx
1252: 260\,\mtot$. In contrast to galactic nuclei, we expect only $\approx
1253: 10\%$ of all binary GW sources {\new in star clusters} to have
1254: eccentricities higher than $0.3$ when they enter the LIGO band, and
1255: $\approx 8\%$ to form within the LIGO band. Binaries which merge due
1256: to 3 or 4--body interactions, are even less likely to have such a high
1257: eccentricity \citep{2006ApJ...640..156G}. \citet{2006ApJ...637..937O}
1258: calculated the expected distribution of eccentricities for binaries
1259: that form in dense star clusters (see their Fig.~3). In their
1260: simulations which had over $1000$ mergers from random encounters, they had no cases that had an eccentricity $>0.3$
1261: {\new when the binary's GWs entered} the LIGO band, and only 3 cases
1262: ($\approx 10^{-3}$) with eccentricity $>0.1$. Their calculations also
1263: included secular effects which could result in mergers with higher
1264: eccentricity \citep{2003ApJ...598..419W}. However, in these cases, the
1265: binary must have merged after one Kozai cycle in a hierarchical
1266: triple. {\new Since the Kozai cycle is a dynamical effect, it operates
1267: on a timescale much shorter than the disruption timescale of the
1268: cluster. Thus, any merger due to the Kozai effect would indicate an
1269: active cluster of BHs must still exist. This is in contrast to a
1270: delayed merger from an ejected binary, which takes of order a Hubble
1271: time to merge. } We discuss other aspects of GW detection in
1272: \S~\ref{sec:detection} below.
1273:
1274:
1275: In most instances, the first binary forming encounters occur
1276: within the LIGO band. However, the S/N of such encounters is too small
1277: to be detected for low masses \citep{2006ApJ...648..411K}. Therefore,
1278: encounters which remain eccentric throughout the inspiral, especially
1279: near plunge, may be the most readily detectable encounters. Therefore
1280: we are interested in the eccentricity of the binary as it reaches the
1281: LSO (see Eq.~\ref{eq:LSO}). From Eq.~(\ref{eq:dgdrp}), and the
1282: equation of evolution \citep[][ Eq.~\ref{e:t-e}]{1964PhRv..136.1224P}
1283: we solve for the probability distribution of eccentricity at the LSO.
1284: We plot the eccentricity distribution at LSO in Figure~\ref{fig:dpdlne}
1285: for both galactic nuclei and globular clusters.
1286: For encounters which that
1287: to direct plunge, our calculation gives an eccentricity greater than 1.
1288: However, for normalisation purposes, we include this in our
1289: calculations, as they interestingly comprise a significant fraction of
1290: merger events.
1291:
1292: \begin{figure}
1293: \begin{center}
1294: \includegraphics[width=\columnwidth]{erpabw.ps}
1295: \end{center}
1296: \caption{\label{fig:inspirala} The secular evolution of BH binaries.
1297: We have plotted in $10\%$ probability intervals the evolution of the
1298: binaries as they decay ($M = m = 10\,\Msun$; see
1299: Fig.~\ref{fig:distribution} for details). The solid line is the
1300: orbit averaged evolution given by
1301: Eqs.~(\ref{e:rho_p1})~\&~(\ref{e:rho_p2}). The solid circles denote
1302: the completion of approximately one orbital period, $\Delta t_{\rm
1303: orb}$. The gap between the solid line and $e=1$ is due to the
1304: finite loss of energy during the initial parabolic encounter. For
1305: nearly $\sim 30\%$ of all binaries, the orbit averaged approximation
1306: is (visibly) not valid, as can be seen by the large space between
1307: each orbit. The dotted (red) line denotes where the binaries peak harmonic
1308: is $10\,$Hz, the lower limit of the LIGO band. Nearly $90\%$ of all
1309: binaries that form in galactic nuclei are within this limit upon
1310: first passage. The dashed (blue) line denotes the eccentricity at the last
1311: stable orbit (Eq.~\ref{eq:LSO}). }
1312: \end{figure}
1313:
1314:
1315: Until now, nearly all LIGO sources were expected to have a negligible
1316: eccentricity as they enter the LIGO band \citep[but see][for intermediate mass ratio inspirals in star clusters]{2008ApJ...681.1431M}.
1317: The comparatively low eccentricity binary formed through few-body
1318: encounters or standard binary evolution
1319: circularise before they enter the LIGO band and are detected.
1320: Therefore, the detection of eccentric inspirals is a strong test of the
1321: formation scenario of nuclear binaries, and can conclusively reveal the
1322: origin of the BHs.
1323:
1324:
1325: \begin{figure}
1326: \begin{center}
1327: \includegraphics[width=\columnwidth]{dpdlne.ps}
1328: \end{center}
1329: \caption{\label{fig:dpdlne} The eccentricity distribution of events.
1330: Plotted is the eccentricity distribution of mergers at the last
1331: stable orbit ($\rmd \Gamma / \rmd \ln{e_{\rm LSO}}$) for one example
1332: of a galactic nucleus (solid line) and a globular cluster (dotted
1333: line). Both lines are normalised so that they reach a maximum
1334: value of $1$. $e_{\rm LSO} \ge1 $ corresponds to encounters that
1335: directly undergo a plunge. See the text and
1336: Fig.~\ref{fig:distribution} for our assumptions and details of the
1337: calculation.}
1338: \end{figure}
1339:
1340:
1341: \subsubsection{Radius Dependence inside the Galactic Nucleus}
1342:
1343:
1344: In Figure~\ref{fig:binformrate}, we plot the cumulative binary formation
1345: rate for radii larger than $r$, $\Gamma(>r)$, as well as $\rmd \Gamma /
1346: \rmd \ln{r}$. For most models, the total differential rate of binary formation
1347: per logarithmic bin is roughly flat.
1348: Thus, each logarithmic radius interval contributes equally to the rate.
1349: We therefore conclude that the rates
1350: determined are rather robust to the depletion of BHs very close to the
1351: SMBH {\new as may be caused by resonant relaxation \citep{1996NewA....1..149R, 1998MNRAS.299.1231R,2006ApJ...645.1152H}}
1352: or our choice of the innermost radius for BHs. In order for the rate to be
1353: dominated by mergers at large $r$, the number density of the BHs would have
1354: to decrease with an exponent $r^{-\alpha}$ where $\alpha = p + 3/2 <3/2$
1355: (see \S~\ref{sec:massseg}).
1356: This is precisely the reason we accounted for the stars in determining
1357: the potential in
1358: Eq. (\ref{eq:numdens}),
1359: and did not let the density profile of the BHs and
1360: stars go to a constant value as in previous analyses.
1361:
1362: We do not expect these tight binaries in their subsequent inspiral phase to
1363: have any observable effect dynamically. Overall, we expect $\sim 10 -
1364: 10^3$ such binaries to merge over a Hubble time. This presents a much
1365: smaller source of energy than the SMBH, which accretes $\sim 10^4$ BHs
1366: over a Hubble time (FAK06). However small the intrinsic rate in each
1367: galaxy, the cumulative merger rate of many galaxies is large enough to be
1368: detected by future ground based gravitational wave observatories.
1369:
1370:
1371:
1372: \subsubsection{Cosmological Merger Rate Density}
1373: \label{sec:cosmergerrate}
1374:
1375: Using the $\msmbh-\sigma_*$ relationship found by \citet{2002ApJ...574..740T}
1376: for higher mass BHs,
1377: \begin{equation}
1378: \label{eq:msigma}
1379: \msmbh \approx 1.3\times 10^8\,\Msun ( \sigma_* /
1380: 200\,\kms)^4,
1381: \end{equation}
1382: we can extrapolate our results from \S~\ref{sec:binform} to a range of
1383: galactic nuclei and determine the overall rate of mergers in the
1384: Universe. {\new Observations by \citet{2005ApJ...619L.151B} \&
1385: \citet{2006ApJ...641L..21G} have demonstrated that
1386: Eq.~(\ref{eq:msigma}) extends to active SMBHs with masses as low as
1387: $10^5\,\msun$, well within the range of our interest. } We expect
1388: the total mass of stars within the radius of influence to be
1389: comparable to the mass of the SMBH, $M(< r_i) \approx 2 \msmbh$. If
1390: these stars follow a radial density profile of $n_* \propto r^{-3/2}$
1391: as expected for a relaxed system (see \S~\ref{sec:massseg}), then
1392: their number density at the radius of influence is
1393: \begin{equation}
1394: \label{eq:nstar}
1395: n_*(\msmbh) \approx
1396: 1.2\times10^5\,{\rm pc}^{-3} \sqrt{10^6\,\Msun/\msmbh}.
1397: \end{equation}
1398: This gives a number density about $45\%$ larger than assumed in
1399: \S~\ref{sec:massseg} for the Milky Way {\new and used in our calculates presented in Table~\ref{table1}}, but is well within the
1400: expected scatter in number densities as we discuss later in this
1401: section. {\new The actual number density at the radius of influence
1402: for a BH will depend on both the formation history of the galaxy as
1403: well as the merger history of the SMBH. } Because the merger rate is
1404: usually greatest at the smallest radii, we can can determine the rate
1405: in any nucleus by scaling Eq.~(\ref{eq:intsimpfin}) and evaluating the
1406: integral at $r_{\rm min} \propto \sqrt{\msmbh}$, the radius where the
1407: merger timescale of the BH into the SMBH is approximately a Hubble
1408: timescale (approximately the inner radius of BHs). Coincidentally,
1409: this radius has the same scaling relation to the outer radius, $r_{\rm
1410: max} = r_i \propto \sqrt{\msmbh}$ {\new where we use
1411: Eq.~(\ref{eq:msigma}).} The total merger rate, $\Gamma$ is simply
1412: proportional to $\propto n_{\rm BH}^2 (b^2 w) r^3$ evaluated at $r =
1413: r_{\rm min}$. We can approximate the number density of BHs as $n_{\rm
1414: BH}(r) \approx C_{\rm BH} n_*(r_i) (r/r_i)^{-p-3/2}$. Substituting
1415: in $r_i$, we get $n_{\rm BH}{\new (r_i)} \propto \msmbh^{-1/2}$,
1416: independent of the power-law distribution of BHs, $p$. The
1417: cross-section of binary capture times the relative velocity is $\pi
1418: b^2 w \propto w^{-11/7} \propto \msmbh^{-11/28}$, given $w \propto
1419: (\msmbh/r)^{1/2}$. Combining these dependencies, we {\new find the
1420: merger rate has a} relatively weak dependence on the mass of the
1421: SMBH, $\Gamma \propto \msmbh^{3/28}$. Over two orders of magnitude in
1422: mass, we expect the rate to change only by $\approx 40-60\%$. To test
1423: this {\new scaling} relationship we have run a simulation with $\msmbh = 10^5\,\Msun$
1424: and $\sigma_* = 30\,\kms$, and found that the rate was in fact
1425: comparable to the relationship found here. Any discrepancy is likely
1426: due to the slight difference in the capture rate of the BHs and stars
1427: (see Eq.~[\ref{eq:losscone}]).
1428:
1429:
1430: \begin{figure*}
1431: \centering
1432: \includegraphics[width=\columnwidth]{rateB.ps}
1433: \includegraphics[width=\columnwidth]{rateE-2.ps}
1434: \caption{Rate of binary formation as a function of the log of the
1435: radius for Models B and E-2. The solid line is the cumulative rate
1436: of binary formation for radii greater than $r$. The dotted line
1437: is the differential rate distribution per logarithmic bin ($\rmd
1438: \Gamma / \rmd\log{r}$).}
1439: \label{fig:binformrate}
1440: \end{figure*}
1441:
1442:
1443: Although the merger rate of BHs is not very sensitive to the SMBH
1444: mass, it is sensitive to the intrinsic scatter of $n_*$ {\new for each
1445: galaxy}, which is used to normalise the entire distribution of of
1446: the BHs {\new (see Eq.~[\ref{eq:numdens}])}. The expected rate of mergers is determined by $\langle
1447: n_*^2\rangle$, the average of the number densities squared at the
1448: radius of influence {\new for the distribution of galaxies in the
1449: universe}. Given $\langle n_*^2\rangle = \langle n_*\rangle^2 +
1450: \sigma_n^2$, where $\sigma_n$ is the variance of the number density at
1451: the radius of influence {\new of a population of SMBHs}, we must
1452: rescale our results by a factor
1453: \begin{equation}
1454: \label{eq:xi}
1455: \xi = \frac{\langle n_*^2\rangle}{\langle n_*\rangle^2} = 1+\frac{\sigma_n^2}{\langle n_*\rangle^2}.
1456: \end{equation}
1457: {\new Previous studies on GW event rate estimates neglected corrections
1458: due to cosmic variance, Eq.~(\ref{eq:xi}). However it is quite plausible
1459: that {\it not} all galaxies with the same SMBH mass have exactly the same
1460: number of stars in the central cusp. In fact,}
1461: \citet{2007ApJ...671...53M} determined the relaxation times for
1462: galaxies in the ACS Virgo Survey \citep{2004ApJS..153..223C}, and
1463: found a relatively tight correlation, however, there was still
1464: significant scatter {\new above} the mean by about an order of
1465: magnitude. We have taken the results from \citet{2007ApJ...671...53M}
1466: (specifically all nuclei with $\sigma_* < 140\,\kms$ in their Fig.~1),
1467: and determined that $ \sigma_n^2/\langle n_*\rangle^2 \sim 30$. {\new
1468: Although there is considerable uncertainty in the actual value of
1469: $\xi$, both observationally and theoretically, we expect the merger
1470: rate to be larger than we have so far calculated by $\xi\gtrsim
1471: 10$--$100$, and scaled our results by $\xi_{30} = \xi / 30$. In our
1472: calculations, however, we use the conservative slope of
1473: Eq.~(\ref{eq:nstar}) to determine the mean value of $n_*$, and not
1474: the shallower slope found by \citet{2007ApJ...671...53M}, which
1475: would give rate estimates to be an order of magnitude larger than
1476: found here.}
1477:
1478:
1479: We calculate the average cosmological merger rate density by
1480: convolving the rate per galaxy with the number density distribution of
1481: SMBHs in the universe. We extrapolate the results of
1482: \citet{2002AJ....124.3035A}, who found the best fit
1483: number density distribution of massive SMBHs to be
1484: \begin{equation}
1485: \label{eq:smbhnumber}
1486: \frac{\rmd n_{\rm SMBH}}{\rmd \msmbh} = c_o\left(\frac{\msmbh}{M_\bullet}\right)^{-1.25} e^{-\msmbh /M_\bullet},
1487: \end{equation}
1488: assuming this formula is valid all the way to $\msmbh=10^{4}\,\Msun$,
1489: where $c_o = 3.2\times 10^{-11}\,\Msun^{-1}\,{\rm Mpc}^{-3}$ and
1490: $M_\bullet = 1.3\times 10^8\,\Msun$. Finally, we get the cosmological
1491: merger rate by integration,
1492: \begin{eqnarray}
1493: \label{eq:cosrate}
1494: {\cal R} &=& \int\limits_{10^4\,\Msun}^{10^7\,\Msun} \Gamma_{1\rm GN}(\msmbh) \frac{\rmd n_{\rm SMBH}}{\rmd
1495: \msmbh} \rmd \msmbh \nonumber\\&\approx& 3\, \Gamma_{1\rm GN} \xi_{30}\,{\rm Mpc}^{-3},
1496: \end{eqnarray}
1497: where $\Gamma_{1\rm GN}$ is the expected rate of mergers for a single
1498: galactic nucleus of a specific model shown in Table~\ref{table1}. The
1499: normalisation $3\xi_{30}\,{\rm Mpc}^{-3}$ follows from the
1500: distribution given in Eq.~(\ref{eq:smbhnumber}), and also accounts for
1501: the intrinsic scatter of $n_{*}$ {\new for a population of galaxies},
1502: Eq.~(\ref{eq:xi}). For our fiducial model of the MW, Model B, we get
1503: a comoving rate density of $8.4\times 10^{-10} \xi_{30}\,{\rm
1504: yr}^{-1}\,{\rm Mpc}^{-3}$. Because this integral is nearly flat in
1505: the log of $\msmbh$, it results in a similar rate density per
1506: logarithmic mass bin and {\new is not very sensitive to the limits of
1507: the integration. Taking the lower limit of currently observed SMBH
1508: masses, of $10^5\,\msun$, we get a rate $\approx 60\%$ of what we
1509: calculated here. However, recent observations by \citet{2008ApJ...688..159G} have shown that the $\msmbh-\sigma$ relation extends to
1510: the smallest SMBHs observed, some of which reside in galaxies
1511: without a classical bulge. Hence, } there is still $\approx 50\%$
1512: uncertainty in the merger rate due to the true function of
1513: $\Gamma_{1\rm GN}(\msmbh)$ discussed above, but the total rate density
1514: is relatively robust given the other uncertainties in our calculation,
1515: especially $\xi_{\rm 30}$ and $C_{\rm BH}$.
1516:
1517: For future calculations, we define $\rmd {\cal R}_{mM}/\rmd
1518: {\rho_{p0}}$ as the rate density for fixed masses $m$ and $M$
1519: analogous to Eq.~(\ref{eq:diffsimpfin}) and (\ref{eq:dgdrp}).
1520:
1521: \subsubsection{Application to Massive Star Clusters}
1522: \label{sec:clusters}
1523:
1524: BHs in massive star clusters {\new without a central massive black
1525: hole} will also undergo an epoch of mass segregation
1526: \citep{1993Natur.364..421K,1993Natur.364..423S}, in which the BHs
1527: segregate to the cluster core, and effectively decouple from the stars
1528: forming their own subcluster
1529: \citep{2000ApJ...528L..17P,2004ApJ...608L..25M,2006ApJ...637..937O,2008arXiv0804.2783M}.
1530: This BH subcluster will continue to interact only with the BHs until a
1531: sufficient number of BHs are ejected dynamically, that they come back
1532: into equilibrium with the stars. This occurs approximately when
1533: $N_{\rm BH} \lesssim 100$ \citep{2000ApJ...539..331W}. For massive
1534: clusters, with $\langle w_{\rm BH}\rangle \approx 15\,\kms$ and
1535: $n_{\rm BH} \sim 10^6\,$pc$^{-3}$, the cluster evaporates before about
1536: $\lesssim 10^8-10^9\,$yr. Scaling Eq.~(\ref{eq:intsimpfin}) to these
1537: parameters, the cluster will have a BH-BH merger rate of
1538: \begin{eqnarray}
1539: \Gamma_{1\rm MSC} &\approx& N_{\rm BH} \left< n_{\rm BH} w_{\rm BH} \sigma_{cs} \right> \nonumber\\&\approx&
1540: 1.8\times 10^{-8}\,{\rm yr}^{-1} \times\left(\frac{N_{\rm BH}}{1000}\right)
1541: \left(\frac{n_{\rm BH}}{10^6\,{\rm pc}^{-3}}\right)\nonumber\\
1542: &&\quad \times\left(\frac{w_{\rm BH}}{15\,\kms}\right)^{-11/7},
1543: \label{eq:clusterrate}
1544: \end{eqnarray}
1545: during this period of evolution, {\new if the number density of BHs is
1546: uniform in radius}.
1547: This is comparable to, but slightly less than merger the rate found by
1548: \citet{2006ApJ...637..937O} for the early evolution of a cluster of
1549: BHs due to three-body and four-body encounters alone. The detection
1550: rate of such early mergers depends on the number of {\em young}
1551: clusters (with $t_{\rm age} \lesssim 10^8-10^9\,$yr) within the
1552: detection limit of LIGO. Globular clusters, an important source of
1553: delayed mergers, are too old to still have a BH subcluster. However,
1554: the young clusters in star burst galaxies would be an excellent source
1555: if they survive sufficiently long in their hosts to undergo this
1556: process of mass segregation \citep{2007PhRvD..76f1504O}.
1557:
1558: {\new The eccentricity distribution of binary capture mergers in star
1559: clusters is plotted in Figure~\ref{fig:dpdlne}. Overall, the rate
1560: of mergers in star clusters is dominated by small eccentricity
1561: events, which would be detected as circular inspirals by
1562: ground-based gravitational wave detectors. However,} young star
1563: clusters may have as many, or even more, eccentric mergers than are
1564: expected in the nuclei of galaxies {\new if there are a sufficient
1565: number of mergers in young star clusters}. In massive star
1566: clusters, we expect $\sim 10\%$ of all gravitational wave captures to
1567: merge with eccentricities similar to those in galactic
1568: nuclei. Therefore, the distribution of low eccentricity events will be
1569: indicative of the source of BH-BH mergers, and may be useful in
1570: constraining the distribution and evolution of BHs in both galactic
1571: nuclei and massive star clusters.
1572:
1573: \section{Detection of gravitational waves}
1574: \label{sec:detection}
1575:
1576: To determine the expected detection rate of sources, we must now
1577: calculate the maximum luminosity distance to which these inspirals are
1578: detectable. In this section, we discuss the general properties of the
1579: waveform, calculate the maximum distance of detection, and add up the
1580: total expected detection rate for second generation terrestrial GW
1581: instruments.
1582:
1583: \subsection{General Properties of the Inspiral}
1584: The evolution of the binary and the GW signal can
1585: be separated into three phases:
1586:
1587: \noindent
1588: {\bf [I]} Highly eccentric encounters -- train of distinct GW bursts
1589: in time, broadband signal in frequency.
1590:
1591: \noindent
1592: {\bf [II]} Moderate-small eccentricity inspiral -- continuous GW
1593: signal in time, dominated by distinct frequency harmonics.
1594:
1595: \noindent
1596: {\bf [III]} Merger and ringdown -- short duration peak GW power and
1597: exponential decay.
1598:
1599: \begin{figure}
1600: \centering
1601: \mbox{\includegraphics{time_orbtime_ptime.eps}}
1602: \caption{\label{f:t-dt} The relevant timescales that determine the GW
1603: waveform: the evolution of the orbital time ($\Delta t_{\rm orb}$) and the
1604: time duration of pericenter passage ($\Delta t_p$) as functions of time to
1605: merger ($t_{\rm merger}$). The units are minutes on both axes for
1606: $(M,\eta,\rho_{p0})=(20,.25,40)$ and are different for other values as
1607: marked. Similar to Fig.~\ref{f:t-e} the orbits should not be extrapolated
1608: beyond the LSO for a particular $\rho_{p0}$, shown with circles. The
1609: inspiral is quasiperiodic if $\Delta t_{\rm orb}\ll t_{\rm merger}$ and
1610: $\Delta t_p\ll t_{\rm merger}$. The GW signal is burstlike if
1611: $\Delta t_p \ll \Delta t_{\rm orb}$ and is continuous if $\Delta t_p \sim
1612: \Delta t_{\rm orb}$.}
1613: \end{figure}
1614:
1615: The distinction between Phases {\bf I} and {\bf II} can be understood
1616: by studying the evolution of the relevant timescales that determine
1617: the GW waveform: the orbital time ($\Delta t_{\rm orb}=\omega_{\rm
1618: orb}^{-1}$), and the time duration of pericenter passage ($\Delta
1619: t_p=\omega_{p}^{-1}$). These are plotted in Figure~\ref{f:t-dt} as a
1620: function of time to merger $t_{\rm mg}$. For both Phases~{\bf I} and
1621: {\bf II}, $|\rmd \Delta t_{\rm orb}/\rmd t_{\rm mg}|\ll 1$, so the
1622: orbit is quasiperiodic and evolves gradually due to the emission of
1623: GWs.
1624: %
1625: The characteristic GW emission timescale during each orbit is determined by
1626: $\Delta t_p$. Figure~\ref{f:t-dt} shows that initially (Phase~{\bf I}), $\Delta t_p\ll \Delta t_{\rm orb}$, implying
1627: that the waveform consists of a train of short $\Delta t_p$ duration bursts arriving quasiperiodically
1628: with separation $\Delta t_{\rm orb}$. Later, when the burst duration timescale $\Delta t_p$ becomes comparable to
1629: $\Delta t_{\rm orb}$, the signal becomes continuous in time domain (Phase~{\bf II}).
1630: Since the orbital evolution is quasiperiodic, the GW signal
1631: is approximately a sum of discrete frequency harmonics of the orbital frequency, $f_{\rm
1632: orb}$. When the eccentricity becomes relatively small $e \lesssim
1633: 0.7$, (Phase~{\bf II}) the harmonic decomposition is quickly convergent, while
1634: during Phase~{\bf I}, it is more convenient to work with the continuous limit
1635: of the frequency spectrum.
1636:
1637: During Phase~{\bf I} and {\bf II} the equations of motion and the GW
1638: waveforms can be calculated {\new accurately in
1639: various approximations (see \S~\ref{sec:motivation} above)}.
1640: {\new For Phase~{\bf II}, the expected signal to noise ratio of detection
1641: has been calculated by \citet{\bc}, and the $e=1$ parabolic case was
1642: examined in \citet{2006ApJ...648..411K}. In the following, we
1643: generalise these studies to be applicable to Phases~{\bf I}
1644: and {\bf II}.}
1645:
1646: Once the orbital separation reaches the last stable orbit, the BHs
1647: fall in rapidly and form a common horizon. The last stable orbit,
1648: which marks the end of Phase~{\bf II}, is determined by the initial
1649: pericenter distance $\rho_{p0}$ for an initially parabolic orbit
1650: shown by circles in Figure~\ref{f:t-dt}. The GW waveform during
1651: Phase~{\bf III} involves the calculation of the violently changing
1652: spacetime and the eventual relaxation into a Kerr BH. This requires
1653: full numerical simulations of the Einstein equations. The
1654: detectability of the resulting waveforms have been examined for
1655: nonspinning binaries and quasicircular initial conditions
1656: \citep{\baker,\berti}. These studies have shown that $S/N=10$ can be
1657: reached up to a distance $d_L=1$ -- $6\,$Gpc for total binary mass
1658: $\mtot = 10$--$200\Msun$ for AdLIGO. Eccentric mergers were
1659: considered very recently by \citet{\hinderb,\hindera} and
1660: \citet{\w} for the case of no initial spins, who showed that the
1661: resulting GW power is comparable to (or sometimes larger than) the
1662: power released during quasicircular mergers.
1663: Future studies should address spin effects during the coalescence,
1664: they might significantly modify the GW
1665: power and waveforms.
1666:
1667: We note that the separation between Phases {\bf I}-{\bf II}-{\bf III}
1668: is valid only if the initial encounter has a minimum separation,
1669: $r_p$, that is much larger than the unstable circular orbit, $r_{\rm
1670: UCO} \sim 2$--$4\mtot$, depending on BH spins. Direct captures, or
1671: orbits outside but repeatedly approaching the unstable circular orbit
1672: (the so-called `zoom-whirl' orbits), are qualitatively
1673: different. Such encounters have been studied in the geodesic
1674: approximation appropriate for extreme mass ratios
1675: \citep{\gairof,\gairos,\pk,2008PhRvD..77j3005L} and using full numerical simulation for
1676: equal masses \citep{\hinderb}. In this case the GW spectrum is
1677: considerably different and the power is considerably increased.
1678:
1679: The purpose of this section is to derive the signal-to-noise ratio for
1680: the quasiperiodic Phases {\bf I} and {\bf II} and determine the maximum
1681: range of detection for second generation terrestrial GW instruments.
1682: We leave the assessment of Phase~{\bf III} and the zoom-whirl domain to future studies.
1683:
1684: \subsection{Signal to Noise Ratio}
1685:
1686: {\new Here we briefly review the general calculation of the signal to noise ratio for
1687: detecting the GW signal, which we can then utilise for the waveforms generated by
1688: GW capture events. We refer the reader to \citet{1998PhRvD..57.4535F} for more
1689: details.}
1690:
1691: In general, the signal to noise ratio of a GW detection is defined as
1692: \begin{equation}\label{sn}
1693: \frac{S^2}{N^2} = 4 \int_{f_{\min}}^{f_{\max}}\frac{|h^2(f,{\bm \theta})|}{S_h(f)} \D f,
1694: \end{equation}
1695: where $h(f, {\bm \theta})$ is the Fourier transform of the GW signal
1696: weighted by the antenna beam patterns, ${\bm \theta}=\{\theta_{i}\}$ are
1697: the physical parameters describing the source and the detector orientation,
1698: $S_h(f)$ is the one-sided noise spectral density in units of $\Hz^{-1}$,
1699: and $f_{\min}\leq f \leq f_{\max}$ correspond to the frequency band of the
1700: instrument, e.g. $(f_{\min},f_{\max})\approx (10, 10^4)\,$Hz for AdLIGO\footnote{http://www.ligo.caltech.edu/advLIGO/scripts/summary.shtml}.
1701:
1702: The sky position and binary orientation averaged root mean square
1703: signal-to-noise for a single orthogonal arm interferometric GW
1704: instrument is
1705: \begin{equation}\label{sn2}
1706: \left\langle \frac{S^2}{N^2}\right\rangle = \int_{f_{\min}}^{f_{\max}}\frac{h^2_{c}(f)}{5 f S_h(f)}\, \frac{\D f}{f},
1707: \end{equation}
1708: where $h_{c}(f)$ is the characteristic isotropic GW amplitude
1709: defined as
1710: \begin{equation}\label{hc}
1711: h_{c} = \frac{1}{\pi d_L}\sqrt{2 \frac{\D E}{\D f}} = \frac{1}{\pi d_L}\sqrt{\frac{2\dot E}{\dot f}}\, ,
1712: \end{equation}
1713: where $d_L(z)$ is the luminosity distance to a source at a cosmological
1714: redshift $z$, $\D E/\D f$ is the one-sided GW energy spectral density on a
1715: spherical shell at infinity. The second equality corresponds to the
1716: stationary phase approximation for a quasiperiodic signal sharply peaked at
1717: frequency $f$. In this case, $\dot E$ is the total GW power at frequency
1718: $f$, which evolves slowly in time according to its time derivative $\dot f$.
1719:
1720: Equations~(\ref{sn2}) and (\ref{hc}) can be generalised for a signal
1721: consisting of discrete harmonics $f_n$, with negligible overlap as
1722: \citep{\bc}
1723: \begin{equation}\label{sn3}
1724: \left\langle \frac{S^2}{N^2}\right\rangle =
1725: \sum_{n=2}^{\infty}\int_{f_{\min}}^{f_{\max}}\frac{h^2_{c,n}(f_n)}{5 f_n S_h(f_n)}\, \frac{\D f_n}{f_n},
1726: \end{equation}
1727: and
1728: \begin{equation}\label{hcn}
1729: h_{c,n} = \frac{1}{\pi d_L}\sqrt{\frac{2\dot E_n}{\dot f_n}}\, ,
1730: \end{equation}
1731: where $\dot E_n$ is the GW power radiated at frequency $f_n$. For
1732: quasiperiodic orbits with an intrinsic orbital frequency $f_{\rm orb}$, the
1733: observed frequency harmonics at redshift $z$ are given by
1734: \begin{equation}
1735: \label{eq:forb}
1736: f_n \equiv n f_{{\rm orb},z}\equiv n \frac{f_{\rm orb}}{1+z}.
1737: \end{equation}
1738:
1739: \subsection{Application to the GW Capture Process}
1740: \label{sec:luminositydistance}
1741:
1742: Let us now turn to the detectability of the GWs
1743: starting from the initial hyperbolic encounter and ending
1744: in the violent BH merger. {\new Here we
1745: derive a computationally more efficient equivalent form of Eq.~(\ref{sn3}),
1746: which can be utilised for Phases~{\bf I} and {\bf II}.}
1747:
1748: \begin{figure*}
1749: \centering
1750: \mbox{{\includegraphics{SNR_lne_f_M20_rp10-small.eps}}
1751: {\includegraphics{SNR_lne_f_M20_rp20-small.eps}}
1752: {\includegraphics{SNR_lne_f_M20_rp40-small.eps}}}\\
1753: \mbox{{\includegraphics{SNR_lne_f_M60_rp10-small.eps}}
1754: {\includegraphics{SNR_lne_f_M60_rp20-small.eps}}
1755: {\includegraphics{SNR_lne_f_M60_rp40-small.eps}}}
1756: \caption{\label{f:snr-eccentricity-frequency} The frequency evolution of
1757: GWs as a function of eccentricity for various total mass and initial
1758: periapsis as labelled.
1759: The shading represents the expected signal-to-noise ratio of the first
1760: $n=100$ harmonics per logarithmic eccentricity bins for AdLIGO. The
1761: waveform is described by a broadband spectrum at large $e$ (Repeated Burst
1762: Phase) that later separates into distinct harmonics as eccentricity
1763: decreases until the LSO (Eccentric Inspiral Phase). The signal-to-noise
1764: ratio is substantial already at $e\gtrsim 0.7$. }
1765: \end{figure*}
1766:
1767: {\new Following {\citet{\bc}}}, we calculate the binary evolution and the GW signal waveform in the
1768: leading order approximation of \citet{\pmat}, where the interacting
1769: masses move on quasi-Newtonian trajectories and emit quadrupolar
1770: radiation. This approximation is adequate in terms of the angular-averaged
1771: $S/N$ if the initial periapse is well outside the unstable circular orbit, e.g.
1772: $r_p \gg r_{\rm UCO}$. At smaller initial $r_p$, the Newtonian
1773: approximation underestimates the GW power and is therefore
1774: conservative.\footnote{{\new Modulations due to general relativistic pericenter
1775: precession and relativistic beaming would be very important for the real data analysis, but
1776: not in terms of the the calculation of event rates that depend on the angular-averaged total $S/N$.}}
1777: To lowest order,
1778: \begin{equation}\label{dotEn}
1779: \dot E_n = \frac{32}{5} \eta^2 \mtot^{10/3}\omega_{\rm orb}^{10/3} g(n,e),
1780: \end{equation}
1781: where $g(n,e)$ determines the relative power of the $n$th harmonic for
1782: orbits with eccentricity $e$, which is given by \citet{\pmat} in the
1783: Newtonian approximation as
1784: \begin{eqnarray}\label{gne}
1785: g(n,e) &=& \frac{n^4}{32} \left\{\left[ J_{n-2} - 2eJ_{n-1} + \frac{2}{n}J_{n}
1786: + 2e J_{n+1} - J_{n+2} \right]^2 \right.\nonumber \\
1787: &&\left.
1788: + (1-e^2)[ J_{n-2}- 2J_{n} + J_{n+2}]^2 + \frac{4}{3n^2}J_{n}^2
1789: \right\}.
1790: \end{eqnarray}
1791: Here $J_i(x)$ is the $i$th Bessel function, and we have suppressed the argument
1792: $x=n e$, i.e. $J_i \equiv J_i(ne)$ for each $i$ above. These waveforms generally
1793: have a maximum at frequency $\omega_p$ for $e\gsim 0.5$ and have a steep cutoff
1794: such that the fractional GW power beyond $5\omega_p$ is smaller than $10^{-3}$
1795: \citep{\tu}. For smaller $e$, the harmonics beyond the first $2\omega_p$ are
1796: greatly suppressed. The number of harmonics necessary to a precision $10^{-3}$
1797: can be estimated as
1798: \begin{equation}\label{nmax}
1799: n_{\max} = 5 \frac{\omega_{p}}{\omega_{\rm orb}} = 5 \frac{(1+e)^{1/2}}{(1-e)^{3/2}}
1800: \end{equation}
1801: e.g. $n_{\max}=(10,40,10^4)$ for $e=(0.3,0.7,0.99)$, respectively.
1802:
1803: Note that Eq.~(\ref{dotEn}) depends explicitly on $e$. Therefore in
1804: order to directly evaluate the $S/N$ integral in Eq.~(\ref{sn3}) over
1805: $f_n$ one needs to invert the frequency evolution equation
1806: $f_n(r_{p0},e_0,e)$. Computationally it is much more efficient to
1807: change the integration variable from $f_n$ to $e$,
1808: \begin{equation}
1809: \frac{\D f_n}{f_n} = - \frac{3}{2}\D\ln \frac{a}{a_0} =
1810: -\frac{18}{19}\frac{1 + \frac{73}{24} e^2 + \frac{37}{96} e^4}{
1811: 1 - \frac{183}{304}e^2 - \frac{121}{304}e^4}
1812: \frac{\D e}{e},
1813: \end{equation}
1814: and reverse the order of the sum and the integral
1815: \begin{equation}
1816: \left\langle\frac{S^2}{N^2}\right\rangle =
1817: \frac{48}{95} \frac{\eta M_{{\rm tot},z}^3 \rho_{p0}^{2}}{d_L^2}
1818: \int\limits_{e_{\rm LSO}}^{e_{0}}\sum_{n=2}^{n_{\max}(e)}
1819: \frac{g(n,e)s(e,e_0)}{n^2 S_h(f_n)}
1820: \frac{\D e}{e},\label{sn4-I}
1821: \end{equation}
1822: where $e_{\rm LSO}$ corresponds to the particular $\rho_{p0}$ (see Eq.~\ref{eq:LSO}),
1823: $e_0$ is given by Eq.~(\ref{eq:e}), and
1824: \begin{equation}
1825: s(e,e_0)=\left(\frac{e}{e_0}\right)^{\frac{24}{19}}
1826: \left(\frac{1+\frac{121}{304}e^2}{1+\frac{121}{304}e_0^2}\right)^{\frac{1740}{2299}} \frac{(1+e_0^2)(1-e^2)^{3/2}}{1-\frac{183}{304} e^2 -\frac{121}{304} e^4}.
1827: \end{equation}
1828: Note, that $S_h(f)=\infty$ is assumed outside of $f_{\min}\leq f\leq f_{\max}$.
1829: The upper limit of the sum in this form can be adjusted to the required calculation precision using Eq.~(\ref{nmax}).
1830:
1831: For Phase~{\bf II}, we can rewrite Eq.~(\ref{sn4-I}) by changing the sum
1832: over $n$ to a continuous integral over $f_n$.
1833: \begin{equation}
1834: \left\langle\frac{S^2}{N^2}\right\rangle =
1835: \frac{48}{95}\frac{\eta M_{{\rm tot},z}^3 \rho_{p0}^{2}}{d_L^2}
1836: \int\limits_{e_{\rm LSO}}^{e_0}
1837: \int\limits_{f_{\min}}^{f_{\max}}
1838: \frac{g(n,e)s(e,e_0)}{{f_{\rm orb},z} S_h(f)}
1839: \frac{\D f}{f}\frac{\D e}{e}.\label{sn4-II}
1840: \end{equation}
1841: where $n=f/f_{{\rm orb},z}$, $f_{{\rm orb},z}=f_{\rm orb}(r_{p,z},e,e_0)$
1842: given above by Eq.~(\ref{eq:forb}), and $r_{p,z} = M_{{\rm tot},z} \rho_{p0}=(1+z)M_{\rm tot} \rho_{p0}$.
1843:
1844:
1845: In addition to their numerical advantages, Eqs.~(\ref{sn4-I}) and
1846: (\ref{sn4-II}) can be used to study the time-frequency evolution of the
1847: instantaneous $S/N$ accumulation rate as the orbit evolves. Figure~\ref{f:snr-eccentricity-frequency}
1848: shows the contribution of the first $n\leq 100$ harmonics to the
1849: signal-to-noise ratio for AdLIGO (i.e. before evaluating the sum or the
1850: integral in Eq.~\ref{sn4-I}) for total masses $\mtot=(20,60)\Msun$ and
1851: initial periapse $\rho_{p0}=(10,20,40)$. The figure illustrates the
1852: unique frequency evolution of the signal consistent with the
1853: expectations described above. Initially during Phase~{\bf I}, it is
1854: broadband in frequency, and decouples into discrete harmonics at
1855: smaller eccentricities during Phase~{\bf I}. The contribution of upper
1856: harmonics is nonnegligible even at LSO, especially if the initial
1857: periapse satisfies $\rho_{p0}\lsim 40$. Note that the maximum
1858: frequency of the $S/N$ at large eccentricities $e\gtrsim 0.8$ in
1859: Figure~\ref{f:snr-eccentricity-frequency} is merely a consequence of
1860: not plotting harmonics beyond $n=100$, leading to a large
1861: underestimate of $S/N$ in Phase~{\bf II}. In this case Eq.~(\ref{sn4-II})
1862: becomes more useful than Eq.~(\ref{sn4-I}).
1863:
1864: \begin{figure}
1865: \centering
1866: \mbox{\includegraphics{dSNR_adLIGO_lne.eps}}
1867: \caption{\label{f:dsnr-eccenticity} The square of the
1868: signal-to-noise ratio per
1869: logarithmic eccentricity bin as a function of eccentricity for
1870: AdLIGO at $1\,$Gpc. Three sets of curves are shown for total masses with
1871: different line types as labelled and initial periapse
1872: $\rho_{p0}=(10,20,40)$ from right to left corresponding to
1873: Figure~\ref{f:snr-eccentricity-frequency}. Eccentric encounters
1874: between massive black holes are detectable up to $700\Msun$ with
1875: AdLIGO that are totally invisible during a circular inspiral. }
1876: \end{figure}
1877:
1878: The frequency-independent $S/N$ accumulation rate can be obtained as a
1879: function of eccentricity (or time using the $e(t)$ dependence shown in Fig.~\ref{f:t-e}),
1880: if evaluating the sum over $n$ in
1881: Eq.~(\ref{sn4-I}) or the integral over $\D f/f$ in Eq.~(\ref{sn4-II}),
1882: but not the integral over
1883: eccentricity. Figure~\ref{f:dsnr-eccenticity} shows the result for the
1884: same masses and periapse as Fig.~\ref{f:snr-eccentricity-frequency},
1885: including an additional extreme case of $\mtot=700\Msun$. For BH masses
1886: below $\mtot\sim 100$, the $S/N$ contribution of the high and low
1887: eccentricity phases are comparable. The $S/N$ during highly eccentric
1888: encounters dominate for the typical case $\mtot\sim 20\Msun$ for
1889: $\rho_{p0}\lesssim 20$, while small $e$ dominates for larger masses
1890: $\mtot\sim 60\Msun$ for $\rho_{p0}\gtrsim 20$. It is very
1891: interesting that the eccentric encounters between binaries with intermediate
1892: masses up to $\mtot=700\Msun$ are detectable to $\approx 1\,$Gpc for close encounters
1893: $\rho_{p0}\lesssim 10$ with AdLIGO, because these masses are otherwise
1894: totally invisible during a circular inspiral. This is explained by the
1895: broadband nature of the signal during Phase~{\bf I}, leading to
1896: nonnegligible power leaking into the detectable frequency range of the
1897: detector $f\geq 10\Hz$. Such massive BHs were thought to be accessible
1898: only during the violent merger/ringdown phase for AdLIGO \citep{1998PhRvD..57.4535F,2007PhRvD..75l4024B}.
1899:
1900: The total $S/N$ can be obtained by evaluating both the sum and the
1901: integrals in Eqs.~(\ref{sn4-I}) and (\ref{sn4-II}). The result can be
1902: converted into a maximum distance of detection assuming a detection
1903: threshold, e.g. $\langle S^2/N^2 \rangle\equiv 5^2$. For Phase~{\bf II},
1904: \begin{equation}
1905: \label{eq:luminositydistance}
1906: d_{L}^{\max} =
1907: \sqrt{\frac{48}{95} \frac{\eta M_{{\rm tot},z}^3 \rho_{p0}^{2}}{\left\langle S^2/N^2 \right\rangle}
1908: \int\limits_{e_{\rm LSO}}^{e_{0}}\sum_{n=2}^{n_{\max}(e)}
1909: \frac{g(n,e)s(e,e_0)}{n^2 S_h(f_n)}
1910: \frac{\D e}{e}},
1911: \end{equation}
1912: and similarly for Phase~{\bf I}. Note, that the integral depends on two parameters $\mtot$ and $\rho_{p0}$ and is independent of the mass ratio $\eta$.
1913:
1914: \begin{figure}
1915: \centering
1916: \mbox{\includegraphics{SNR_Gpc_adLIGO_rp0_M.eps}}
1917: \caption{\label{f:snr} The total signal-to-noise ratio contours as a
1918: function of the pericenter separation $\rho_{p0}$ of the first passage and total
1919: binary mass, $M_{{\rm tot},z}$ for AdLIGO for an equal-mass binary at
1920: $1\,$Gpc. Different contours show $\langle S^2/N^2\rangle^{1/2}$
1921: (averaged over the binary orientation and sky position) increasing
1922: in steps of 5 for the solid and 1 for the dotted lines. For other
1923: mass ratios the SNR is reduced by $(4\eta)^{1/2}$. The limit
1924: $\rho_{p0}\rightarrow \infty$ corresponds to circular inspirals in
1925: the AdLIGO band, which result in typically a smaller $S/N$ than
1926: most likely events with $\rho_{p0}\lesssim 40$. Note that different
1927: solid contours correspond to a maximum distance of detection
1928: increasing in steps of Gpc for a detection threshold $S/N=5$. }
1929: \end{figure}
1930:
1931: Figure~(\ref{f:snr}) shows contours of $d_L^{\max}$ for masses $2\Msun
1932: \leq \mtot \leq 1000\Msun$ and initial periapse $4 \leq \rho_{p0}\leq
1933: 100$. For the limit of large $\rho_{p0}$, the signal is circularised
1934: when it arrives in the LIGO band, corresponding to the circular
1935: inspiral. The figure shows that the eccentric inspirals can be
1936: detected just to a slightly larger distance than the circular
1937: inspirals for masses $\mtot\leq 20\Msun$. The difference becomes more
1938: pronounced at much larger masses. A 100$\Msun$--100$\Msun$ eccentric
1939: inspiral is detectable up to $3\times$ farther than a standard
1940: $10\Msun$--$10\Msun$ circular inspiral. For even larger masses,
1941: $d_{L}^{\max}$ decreases, but remains nonnegligible up to
1942: $\mtot=700\Msun$. Currently less is known about the existence of such
1943: intermediate mass black holes in galactic cusps, which makes it very
1944: difficult to make any theoretical estimates on the expected rates of
1945: such encounters. It is possible that AdLIGO will provide the first direct
1946: observational limit on the population of these objects. We explore
1947: this along with other estimates of the detection rate in \S~\ref{sec:detrate}.
1948:
1949:
1950:
1951: \subsection{Detection Rate Estimates}
1952: \label{sec:detrate}
1953:
1954: We set $\left\langle S^2/N^2 \right\rangle^{1/2} \geq 5$ as our
1955: detection threshold for a single GW instrument, and determine the volume averaged maximum
1956: luminosity distance for detection for each pair of masses, $d_{\rm
1957: L}^{\max}(m,M)$ from Eq.~(\ref{eq:luminositydistance}),
1958: assuming a uniform distribution of $r_p$ out to $r_{pmax}$. Thus, the
1959: total detection rate for Advanced LIGO is estimated to be
1960: \begin{eqnarray}
1961: \label{eq:totalrate}
1962: R &=& \int_0^{\rho_{p0}^{\max}} \rmd \rho_{p0} \int\limits_{M_{\min}}^{M_{\max}} \int\limits_{M}^{M_{\max}} \rmd m \rmd M
1963: \int_0^{z_{\max}} \rmd z \frac{\rmd V_{\rm co}}{\rmd z} \nonumber\\
1964: \quad&&\times\frac{1}{1+z}\frac{\partial^3 {\cal R}}{\partial \rho_{p0} \partial m \partial M}
1965: \end{eqnarray}
1966: where $\partial^3 {\cal R}/\partial
1967: \rho_{p0}\partial m \partial M$ is the comoving partial binary
1968: formation rate density between masses $(m,M)$ at initial periapse
1969: $\rho_{p0}$, given by Eqs.~(\ref{eq:diffsimpfin}) and (\ref{eq:cosrate}),
1970: averaged over the distribution of $\msmbh$ and the number
1971: density normalisation $n_{*}$ (see Eq.~\ref{eq:smbhnumber} and
1972: \ref{eq:cosrate}) at redshift $z$, $n_{\rm gal}$ is the comoving
1973: number density of galaxies at redshift $z$, $z_{\max}$ is the maximum
1974: detectable distance corresponding to $d_{L\max}(\rho_{p0},m,M)$, and
1975: $\rmd V_{\rm co}/\rmd z$ is the comoving volume density corresponding
1976: to the given cosmology. In practice, however, we calculate $R$
1977: discreetly and ignore cosmological
1978: effects, which are not relevant for the next generation of GW
1979: instruments,
1980: \begin{eqnarray}
1981: \label{eq:totalrateapprox}
1982: R &\approx& \sum_{M,m<M} \int_0^{\rho_{p0}^{\max}} \rmd \rho_{p0}
1983: \frac{\rmd {\cal R}_{mM}}{\rmd \rho_{p0}}
1984: \frac{4 \pi}{3} (d_L^{\max})^3,
1985: \end{eqnarray}
1986: where $\rmd {\cal R}_{mM}/\rmd \rho_{p0}$ is the differential rate
1987: density of binary formation between mass bins $m$ and $M$ (see
1988: Eqs.~\ref{eq:diffsimpfin} and \ref{eq:cosrate}). The resulting total
1989: detection rates are shown in the last column of Table~1. Note that the
1990: dependence of $d_{L\max}$ on the binary parameters $(\rho_{p0},m,M)$
1991: lead to an observational bias. For example, since $d_{L\max}$ is
1992: relatively larger for $\rho_{p0}\sim 10$, the detection rate of these
1993: encounters is enhanced relative to larger and smaller $\rho_{p0}$,
1994: even though the intrinsic rate of these encounters is independent of
1995: $\rho_{p0}$. We have plotted the differential detection rate as a
1996: function of $\rho_{p0}$ in Figure~\ref{fig:diffrate} for all mergers
1997: as well as each mass bin.
1998:
1999:
2000: Overall, the most massive BHs in galactic nuclei dominate the
2001: detection rate of mergers for AdLIGO. In
2002: Figure~\ref{fig:detectionrate}, we have plotted the distribution of
2003: detectable mergers as a function of radius for the entire population
2004: of BHs, as well as for each mass bin. The clear domination of the
2005: high mass BHs is caused by a combination of three important factors:
2006: {\it (i)} their number density is significantly enhanced by mass
2007: segregation (\S~\ref{sec:massseg}); {\it (ii)} the signal of the event
2008: is much stronger for larger masses (e.g.,
2009: Eqs.~\ref{sn4-I}~\&~\ref{sn4-II}); and {\it (iii)} the cross-section
2010: for binary capture is greater for larger masses (e.g.,
2011: Eq.~\ref{eq:bmax}).
2012:
2013:
2014:
2015: As discussed in \S~\ref{sec:luminositydistance},
2016: we expect the actual inspiral of a BH with an IMBH may be revealed by
2017: AdLIGO if the event is sufficiently eccentric during plunge. Although we
2018: cannot properly account for this in our analysis, we can attempt to compare
2019: the rate to what we have done in this work. For a large mass ratio, $m\gg
2020: M$, the overall cross-section for forming binaries increases roughly as
2021: $b^2 \propto m^{12/7}$. Compared to $10\,\Msun$ BHs, we expect a
2022: $1000\,\Msun$ BH to have a gravitational wave capture event $\sim 3000$
2023: times as often as a single BH. However, the number of IMBHs in the region
2024: is very uncertain. If we take the optimistic number of $\sim 10$ IMBHs in
2025: a single galactic nucleus as a steady state distribution
2026: \citep{2006ApJ...641..319P}, then we expect a comparable total number of
2027: events to $10\Msun$--$10\Msun$ BH-BH inspirals.
2028:
2029: Our analysis in \S~\ref{sec:massseg} suffers from inaccuracies for the
2030: most massive and rarest BHs. Equation~\ref{eq:fokkerplanck} was
2031: derived assuming a constant density core (and constant relaxation
2032: timescale) for large $r$ (BW76), which is clearly violated in most
2033: galactic nuclei. Typically, the total number of BHs in galactic
2034: nuclei decreased with $M_{\max}$. Despite this decrease in number,
2035: we likely cannot extrapolate our calculations to higher mass BHs such
2036: as IMBHs, in order to see what effect they have on flattening the
2037: density profile of stellar mass BHs.
2038:
2039:
2040:
2041:
2042: \begin{figure*}
2043: \centering
2044: \includegraphics[width=\columnwidth]{dpdrpB.ps}
2045: \includegraphics[width=\columnwidth]{dpdrpE-2.ps}
2046: \caption{\label{fig:diffrate} The pericenter distribution for AdLIGO
2047: detections. Plotted is the differential detection rate ($\rmd
2048: R/\rmd \ln{\rho_{p0}}$) as a function of $\rho_{p0}$ for Models B
2049: (left) and E-2 (right). In both figures, the solid line is
2050: differential rate for the all detections, the alternating dotted
2051: and dashed lines are for each individual mass bin M ($\sum_m \rmd
2052: R_{mM}/(2 \rmd \ln{\rho_{p0}})$) in order of decreasing mass from
2053: top to bottom. Note that function plotted here is different than
2054: in Fig.~\ref{fig:distribution}}
2055: \end{figure*}
2056:
2057:
2058: \begin{figure*}
2059: \centering
2060: \includegraphics[width=\columnwidth]{detrateB.ps}
2061: \includegraphics[width=\columnwidth]{detrateE-2.ps}
2062: \caption{\label{fig:detectionrate} The AdLIGO detection rate for
2063: Models B and E-2. The solid line is the total integrated AdLIGO
2064: detection rate at radii larger than $r$, ($R(>r)$), the alternating dotted and dashed lines
2065: are the contribution to the detection rate of each individual mass bin ($\sum_m
2066: R_{mM}(>r)/2$) in order of decreasing mass from top to bottom.}
2067: \end{figure*}
2068:
2069:
2070:
2071: \section{Summary and Discussion}
2072:
2073:
2074: In this paper, we have analysed two separate problems. First we
2075: determined the multi-mass distribution of BHs in galactic nuclei,
2076: which we then used to analyse the detection rate and merger of
2077: gravitational wave capture binaries. We
2078: integrated the time-dependent Fokker-Planck equations for a variety BH
2079: mass distributions until they reached a steady state. We found,
2080: consistent with previous results, that within a relaxation timescale
2081: at the radius of influence of the SMBH, the BHs and stars form
2082: approximately power-law density cusps ($\propto r^{-3/2-p_0 m / M_{\max}}$)
2083: within the radius of influence of the SMBH. Because the BHs
2084: are more massive than the stars, they have steeper density profiles
2085: ($\propto r^{-2}$) and dominate the dynamics and relaxation processes
2086: in the inner $\sim 0.1\,$pc near the SMBH.
2087:
2088: In such dense population environments, the probability of close flybys
2089: between two BHs is nonnegligible.
2090: Using our results for the steady-state distribution of BHs,
2091: we calculated the expected rate and distribution of GW
2092: capture binaries in galactic nuclei. We showed that after forming,
2093: these binaries rapidly inspiral and merge within hours. Unlike other
2094: sources of merging binaries, the BH binaries in galactic nuclei form
2095: with a characteristic GW frequency {\em inside} the AdLIGO
2096: frequency band, and the detectable GW signal duration is much longer
2097: compared to circular inspirals. In addition, for sufficiently small impact parameters,
2098: the waveforms for such eccentric inspirals are broadband, and can be
2099: detected for much larger masses, up to $\mtot \gtrsim 700\,\Msun$.
2100: This exceeds the BH masses previously considered
2101: detectable \citep{2007PhRvL..99t1102B,2008ApJ...681.1431M}, and as
2102: such it opens a new avenue to probe for the existence of a population of
2103: intermediate mass black holes in galactic nuclei.
2104:
2105: Additionally in \S~\ref{sec:clusters}, we also estimated the rate of
2106: GW captures in young, massive star clusters where BHs decouple
2107: from stars to form a subcluster in the centre.
2108: We found that the rate of GW captures in a single cluster may be
2109: intrinsically larger than in a galactic nucleus. Typically, most
2110: binaries circularise before becoming detectable, however a significant
2111: fraction {\new ($\sim 10\%$) }may still merge with residual eccentricity. Given the vastly
2112: different eccentricity distribution of mergers in galactic nuclei and
2113: massive star clusters, the dominate source of eccentric inspirals can
2114: be determined with only a few detections. The total rate of such
2115: events from massive clusters depends on the relatively unconstrained
2116: number of young clusters within the detection limits of AdLIGO.
2117:
2118:
2119: Finally, we analysed the properties of the
2120: GW signal for the eccentric inspiral of comparable mass BHs, starting
2121: from the initial highly eccentric phase up to the final eccentric
2122: inspiral. We found that the maximum luminosity distance that such
2123: events may be detected using a single AdLIGO-type interferometer is
2124: $1.2\,$Gpc for component masses $m=M\sim 10\,\Msun$, and up to
2125: $3.3\,$Gpc for $m=M\sim 50\,\Msun$. We then used this to calculate
2126: the total detection rate of signals, by using a model merger rate
2127: convolved with a realistic population of SMBHs. We found that the
2128: most massive BHs dominate the detection rate of future ground based
2129: gravitational wave detectors with $\sim (10 - 10^3) \xi_{30}$ events
2130: expected per year, where $\xi_{30}$ is a measure of the expectation
2131: value for the square of number densities near the SMBH (see
2132: Eq.~\ref{eq:xi}).
2133: The overall detection rate is sensitive to
2134: the number fraction of BHs as well as the maximum mass of BHs, and so
2135: future observations will be able to constrain both the average star
2136: formation properties and upper mass of BHs in galactic nuclei.
2137:
2138: Overall, the expected AdLIGO detection rate for GW capture binaries in
2139: galactic nuclei is comparable to estimates for other sources of
2140: gravitational waves. BH-BH binaries that form dynamically in massive
2141: star clusters may be detected by AdLIGO $\sim 10-10^4\,$ times per
2142: year, depending on the total fraction of star formation that occurs in
2143: massive, long lived clusters \citep{2000ApJ...528L..17P,
2144: 2004ApJ...616..221G, 2006ApJ...637..937O, 2006ApJ...640..156G,
2145: 2007PhRvD..76f1504O, 2008arXiv0804.2783M}. Estimates for the
2146: population of merging binaries that formed in the early evolution of
2147: present day globular clusters are $\sim 10-100\,$yr$^{-1}$ alone.
2148: Although globular clusters contain only a small fraction of the mass
2149: of stars in the universe, the merger rate of BHs is sufficiently
2150: enhanced by dynamical interactions that the rate is comparable to those
2151: expected from standard binary evolution for all other stars
2152: (\citealt{2002ApJ...572..407B, 2004ApJ...611.1068B,
2153: 2008ApJ...676.1162S}; however see \citealt{2007ApJ...662..504B}).
2154: Merging neutron stars are also a promising source of gravitational
2155: waves, whose progenitors have been directly observed. Using
2156: constraints from the modest population neutron star-neutron star
2157: binaries known in our galaxy,
2158: \citet{2004ApJ...601L.179K,2004ApJ...614L.137K} estimate that $\sim 20
2159: - 600$ events per year may be detected by AdLIGO. We emphasise that
2160: the rates of GW captures calculated in this paper are {\em
2161: independent} of other sources of GWs, and will occur in addition to
2162: other sources. Given the estimates of sources from a variety of
2163: environments, gravitational wave detectors promise to open a new and
2164: interesting window into the physics, dynamics, and evolution of
2165: compact binaries.
2166:
2167: So far, we have not looked at the interaction between the BHs and
2168: other compact objects such as neutron stars (NSs). The NSs are not
2169: expected to segregate significantly in galactic nuclei, however they
2170: are intrinsically more common than BHs. To test their importance we
2171: have performed an additional run of Model E-2 to look at the expected
2172: detection rate of NS-BH mergers. We found that the total detection
2173: rate of such events is about $\sim 1\%$ of the total rate ($\sim
2174: 1\,$yr$^{-1}$), and merits future study. Gravitational waves from
2175: BH-NS inspirals can provide interesting constraints on the equation of
2176: state for NSs \citep{2002PhRvL..89w1102F}, given model sources from
2177: numerical simulations \citep{2006PhRvD..73b4012F,2008ApJ...680.1326R}.
2178: To date, such analyses have focused on the circular inspiral of BHs
2179: with NSs, however recent population synthesis studies suggest the
2180: number of BH-NS binaries that form may be as rare as those expected
2181: here, $\sim 1\,$yr$^{-1}$ \citep{2007ApJ...662..504B}. Given that GW
2182: capture binaries are predominately eccentric throughout inspiral, we
2183: also expect the newly formed BH-NS binaries to be eccentric. However,
2184: with BH-NS binaries, the tidal effects of the BH on the NS can provide
2185: an additional source of energy dissipation in the binary, which can
2186: enrich the GW signal, and cause it to deviate from the point-mass
2187: approximations calculated here. Future numerical simulations with
2188: eccentric encounters must be done in order to test the importance of
2189: eccentricity on the GW signal for BH-NS binaries.
2190:
2191: Another area we neglected is the influence of
2192: binarity on the merger rate of BHs in galactic nuclei. However, we can
2193: estimate their overall importance by looking at what fraction of
2194: binaries are hard enough to survive in galactic nuclei. Given the
2195: distribution of BH binaries in \citet{2004ApJ...611.1068B}, only a
2196: small fraction, $\lesssim 0.1\%$ of the BH-BH binaries are
2197: sufficiently hard (with an orbital period $\lesssim 10\,$days) not to
2198: be disrupted by repeated encounters with stars or BHs. Interestingly,
2199: this fraction is comparable to the fraction of BHs that merge due GW
2200: capture. The properties of the
2201: merger of such binaries have not been explored yet, however, given the
2202: evidence that such encounters produce few eccentric events in massive
2203: star clusters, we expect that they will not have many eccentric
2204: mergers in galactic nuclei. Unfortunately, the methods we have used
2205: here are not suitable for looking at higher order $N$-body
2206: interactions (such as exchanges, or binary-binary scattering);
2207: simulations similar to FAK06 are more adequate for this type of study.
2208: The rates we calculated might be enhanced if high-dispersion
2209: environments of protogalactic cores without a SMBH also exist,
2210: although such systems are old and not long lived
2211: \citep{1987ApJ...321..199Q,1989ApJ...343..725Q,1990ApJ...356..483Q,1993ApJ...418..147L}.
2212:
2213:
2214: {\new
2215:
2216: In order to estimate the detection rate of gravitational wave
2217: capture binaries in galactic nuclei we had to make assumptions that
2218: introduced uncertainty in our calculations. For the sakeof clarity, we wish
2219: to summarise these uncertainties and the order of magnitude effect
2220: they may have on the actual detection rate for AdLIGO. The largest
2221: uncertainty in the rate calculation is the distribution of BHs in
2222: galactic nuclei, especially the number fraction of BHs ($\sim 10^2$)
2223: and their mass distribution ($\sim 10$) in the centres of galaxies.
2224: Observations of our own galactic centre suggest that the star
2225: formation may be top-heavy, and place the rate on the higher end of
2226: that reported here. One effect we neglect in our calculations is
2227: resonant relaxation, which may reduce the rate $(\sim 10)$ by
2228: depleting the central cusp of BHs. Our final rate, however, is in
2229: practice only logarithmically sensitive to the inner edge of the
2230: cusp of BHs, and the densest cusps that dominate our rate will be
2231: least affected by resonant relaxation. The final rate is also nearly
2232: logarithmically dependent on the population of SMBHs in the local
2233: universe (by a factor of at most $\sim 10$),
2234: but is much more sensitive to the distribution
2235: of densities at the radius of influence of the BHs (by a factor of $\sim
2236: 10$--$10^2$). Finally, there is some uncertainty in the detection rate
2237: given our calculation of the $S/N$ of the mergers ($\sim 10$) since
2238: we do not consider corrections for zoom-whirl orbits and
2239: conservatively ignore the contribution of the merger and ringdown
2240: waveforms of the binary to the signal.
2241:
2242: }
2243:
2244: There are many aspects to the merger of GW capture binaries that
2245: require future work, which will be aided by advanced numerical
2246: simulations of the evolution of these binaries. In our calculations
2247: we used only the leading order (Newtonian) formula to calculate the evolution
2248: of the binaries and the GW waveforms, and
2249: do not account for the GW signal during the plunge and merger of
2250: the binary, nor do we consider the contribution of zoom-whirl orbits.
2251: The recent breakthrough of numerical simulations
2252: finally allows the full GW signal to be calculated directly. Eccentric
2253: mergers promise to have a richer signal than that for circular
2254: binaries. Future studies should incorporate the contribution of
2255: the final coalescence to the GW signal, which would increase the
2256: signal strength generated during the inspiral for the same capture event,
2257: and has a potential to further increase the maximum range of detection.
2258:
2259: We also found that a fraction ($\sim 10-20\%$) of mergers will occur
2260: very close to, or within the estimated last stable orbit of BH
2261: binaries. These capture events may be zoom-whirl orbits, which
2262: dramatically increases the GW signal compared to a regular inspiral
2263: \citep{2007CQGra..24...83P}. Such encounters are inherently in the
2264: strong GR regime, and again require a full numerical treatment of
2265: General Relativity to be studied. A precise assessment of these
2266: sources will further increase our estimates of detection rates. Our
2267: estimates for the number of GW capture binaries in galactic nuclei (and
2268: perhaps in massive star clusters) suggest that such events may be
2269: common enough to be directly detected, and may provide the richest
2270: gravitational wave sources in the universe for ground based detectors.
2271:
2272: \section*{Acknowledgements}
2273: We thank Krzysztof Belczynski, Sam Finn, Scott Hughes, Pablo Laguna,
2274: Ilya Mandel, Szabolcs M\'{a}rka, and Coleman Miller for useful
2275: discussions and Andr\'{a}s P\'{a}l for helping with Figure 9.
2276: This work is supported by in part by NASA grant NNX08AL43G, by FQXi, and by
2277: Harvard University funds. BK acknowledges support from OTKA Grant 68228 and
2278: Pol\'anyi Program of the Hungarian National Office for Research and Technology
2279: (NKTH).
2280:
2281:
2282: \bibliography{p}
2283:
2284: \end{document}
2285:
2286: