1: \documentclass[useAMS]{emulateapj}
2:
3: \usepackage{latexsym,graphicx,natbib,apjfonts}
4: \usepackage{amsmath}
5:
6: %\documentclass[12pt,preprint]{aastex}
7:
8:
9: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% Local macros %%%%%%%%%%%%%%%%%%%%%%%%%%%
10: % Style
11: \newcommand\simless{{\thinspace \rlap{\raise 0.5ex\hbox{$\scriptstyle {<}$}}
12: {\lower 0.3ex\hbox{$\scriptstyle {\sim}$}} \thinspace }} %< or of order
13: \newcommand\simgreat{{\thinspace \rlap{\raise 0.5ex\hbox{$\scriptstyle {>}$}}
14: {\lower 0.3ex\hbox{$\scriptstyle {\sim}$}} \thinspace }} %> or of order
15: % Units
16: \newcommand\msun{\, \rm M_\odot}
17: \newcommand\kms{{\, \rm km\,s^{-1}}}
18: % Names
19: \newcommand\mbh{M_{\rm BH}}
20: \newcommand\vbh{V}
21: \newcommand\mg{M_{\rm gal}}
22: \newcommand\rb{r_{\rm b}}
23: \newcommand\re{R_{\rm e}}
24: \newcommand\rhob{\rho_{\rm b}}
25: \newcommand\vk{V_{\rm kick}}
26: \newcommand\ve{V_{\rm esc}}
27: \newcommand\beq{\begin{equation}}
28: \newcommand\eeq{\end{equation}}
29: \newcommand\vb{V_{\rm Brown}}
30: \newcommand\vbb{V^2_{\rm Brown}}
31: \newcommand\tdf{T_{\rm df}}
32: \newcommand\form{F}
33: \newcommand\mdef{M_{\rm def}}
34: \newcommand\mgal{M_{\rm gal}}
35: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
36:
37: \shortauthors{Gualandris and Merritt}
38: \shorttitle{Ejection of Supermassive Black Holes from Galaxy Cores}
39:
40: \begin{document}
41:
42: \title{Ejection of Supermassive Black Holes from Galaxy Cores}
43:
44: \author{Alessia Gualandris and David Merritt}
45:
46: \affil{Center for Computational Relativity and Gravitation,
47: Rochester Institute of Technology,
48: 78 Lomb Memorial Drive, Rochester, NY 14623}
49:
50: \email{alessiag,merritt@astro.rit.edu}
51:
52: \begin{abstract}
53: Recent numerical relativity simulations have shown that the emission
54: of gravitational waves during the merger of two supermassive black
55: holes (SMBHs) delivers a kick to the final hole, with a magnitude as
56: large as $4000\kms$. We study the motion of SMBHs ejected from
57: galaxy cores by such kicks and the effects on the stellar
58: distribution using high-accuracy direct $N$-body simulations.
59: Following the kick, the motion of the SMBH exhibits three distinct
60: phases. (1) The SMBH oscillates with decreasing amplitude, losing
61: energy via dynamical friction each time it passes through the core.
62: Chandrasekhar's theory accurately reproduces the motion of the SMBH
63: in this regime if $2\lesssim\ln\Lambda\lesssim 3$ and if the
64: changing core density is taken into account. (2) When the amplitude
65: of the motion has fallen to roughly the core radius, the SMBH and
66: core begin to exhibit oscillations about their common center of
67: mass. These oscillations decay with a time constant that is at least
68: $10$ times longer than would be predicted by naive application of
69: the dynamical friction formula. During this phase, the SMBH is
70: typically displaced from the peak of stellar density by roughly the
71: core radius. (3) Eventually, the SMBH reaches thermal equilibrium
72: with the stars. We use straightforward scaling arguments to
73: estimate the time for the SMBH's oscillations to damp to the
74: Brownian level in real galaxies and infer times as long as $\sim
75: 1\,\rm Gyr$ in the brightest galaxies. The longevity of the
76: oscillations makes this mechanism competitive with others that have
77: been proposed to explain double or offset nuclei. Ejection of SMBHs
78: also results in a lowered density of stars near the galaxy center;
79: mass deficits as large as five times the SMBH mass are produced for
80: kick velocities near the escape velocity. We compare the $N$-body
81: density profiles with luminosity profiles of early-type galaxies in
82: Virgo and show that even the largest observed cores can be
83: reproduced by the kicks, without the need to postulate
84: ``hypermassive'' binary SMBHs. Implications for displaced AGNs and
85: helical radio structures are discussed.
86: \end{abstract}
87:
88:
89: \keywords{galaxies:nuclei - stellar dynamics}
90:
91: \section{Introduction}
92: \label{sec:intro}
93:
94: The recent breakthroughs in numerical relativity
95: \citep{Pretorius:05,Campanelli:06,Baker:06a} have allowed a number of
96: groups to evolve binary black holes (BHs) to full coalescence. The
97: final inspiral is driven by emission of gravitational waves, and in
98: typical (asymmetric) inspirals, a net impulse is imparted to the
99: system due to anisotropic emission of the waves
100: \citep{Bekenstein:73,Fitchett:84,Favata:04}. Early arguments that the
101: magnitude of the recoil velocity would be modest for non-spinning BHs
102: \citep{Redmount:89} were confirmed by the simulations, which found
103: $\vk\lesssim 200\kms$ in the absence of spins
104: \citep{Baker:06b,Gonzalez:07a,Herrmann:07}. The situation changed
105: dramatically following the first simulations of ``generic'' binaries,
106: i.e., binaries in which the individual BHs were spinning and in which
107: the spins were allowed to have arbitrary orientations
108: \citep{Campanelli:07a}. Kicks as large as $\sim 2000\kms$ have now
109: been confirmed \citep{Campanelli:07b,Gonzalez:07b,Tichy:07}, and
110: simple scaling arguments suggest that the maximum kick velocity would
111: probably increase to $\sim 4000\kms$ in the case of maximally-spinning
112: holes \citep{Campanelli:07b}. The most propitious configuration for
113: the kicks consists of an equal-mass binary in which the individual
114: spin vectors are oppositely aligned and oriented parallel to the
115: orbital plane. The kick amplitude also depends sensitively on the
116: angle between the BH spin vectors and their linear momenta shortly
117: before the plunge \citep{Lousto:07}.
118:
119: Galaxy escape velocities are $\lesssim 3000\kms$ \citep{MMFHH:04},
120: which means that gravitational wave recoil can in principle displace
121: coalescing supermassive black holes (SMBHs) arbitrarily far from
122: galaxy centers, or even eject them completely. The actual
123: distribution of kick velocities is very uncertain, since it depends on
124: the unknown distribution of binary mass ratios and spins, but most
125: kicks are probably $\lesssim 10^3\kms$. A SMBH that is kicked with
126: less than escape velocity will travel some maximum distance from the
127: galaxy center after which its orbit decays due to dynamical friction;
128: most of the energy loss takes place during passages through the galaxy
129: center. Removal of the SMBH from the core has the effect of
130: transferring kinetic energy to the stars and lowering the core density
131: \citep{Redmount:89,MMFHH:04,Boylan:04}. This implies a more gradual
132: return of the SMBH to a zero-velocity state than in a galaxy with
133: fixed density.
134:
135: In fact, however, the SMBH is not expected to ever reach a state
136: of zero kinetic energy.
137: When its energy falls to a value
138:
139: \begin{equation}
140: \frac{1}{2} \mbh V^2 \approx \frac{1}{2} m_{\star} {\rm v}_{\star}^2
141: \label{eq:vbrown}
142: \end{equation}
143:
144: with respect to the galaxy central potential, where $m_\star$ and
145: ${\rm v}_\star$ are a typical stellar mass and velocity respectively,
146: random gravitational perturbations from stars act to accelerate the
147: SMBH as often as they decelerate it. This is the regime of
148: gravitational Brownian motion \citep{Young:77,BW:76,MBL:07}. A
149: natural definition of the ``return time'' of a kicked SMBH is the time
150: required for dynamical friction to reduce the SMBH's mean kinetic
151: energy to the Brownian value. Applying standard expressions for the
152: dynamical friction force leads one to the conclusion that this would
153: occur in a relatively short time, of order a few orbital periods,
154: after dynamical friction has returned the kicked SMBH to the core.
155:
156: The $N$-body simulations presented here were designed to test these
157: expectations by evaluating the return times of kicked SMBHs and by
158: quantifying the induced changes in galaxy structure. These processes
159: can not be studied accurately using classical dynamical friction
160: theory since the SMBH substantially modifies the core as it recoils
161: and falls back. Approximate $N$-body schemes, e.g. tree or grid
162: codes, are also not well suited to the problem since they can not
163: robustly follow both the early (collisionless) and late (collisional)
164: evolution of the SMBH. Large particle numbers are required in
165: order to cleanly separate the collisional and collisionless regimes.
166:
167: These various requirements can currently be met only with parallel,
168: direct $N$-body codes running on special-purpose supercomputers. Our
169: simulations use the $\phi$GRAPE integrator \citep{Harfst:07} as
170: implemented on {\tt gravitySimulator}, a 32-node supercomputer
171: employing GRAPE-6A accelerator boards \citep{GRAPE-6A}.
172:
173: Our findings are surprising in one important respect. After returning
174: to the core, the kicked SMBH exhibits long-lived oscillations with
175: amplitude comparable to the core radius\footnote{A movie showing the
176: oscillations is available at
177: http://ccrg.rit.edu/Research/Publications.php?paper=0708.0771.}.
178: %http://spiegel.cs.rit.edu/\~{}ag/Movies/anim.gif\,.}.
179: These oscillations eventually decay but with a time constant that is
180: at least an order of magnitude longer than would be predicted by a
181: straightforward application of the dynamical friction equation. We
182: demonstrate that the existence, amplitude and damping time of these
183: oscillations are independent of the number $N$ of ``star'' particles
184: used in the simulations, for $N$ up to $2\times 10^6$. The
185: oscillations are similar to those first reported by R. Miller and
186: collaborators \citep{Miller:92,Miller:96} in their pioneering $N$-body
187: studies of the central regions of galaxies. A number of other authors
188: have reported low effective values of the dynamical friction force as
189: it acts on massive objects that inspiral into constant-density cores
190: \citep{Bontekoe:88,Bertin:03,Read:06} or on rotating bars
191: \citep{Weinberg:02,Klypin:03}. Our use of a high-accuracy,
192: direct-summation $N$-body code combined with large particle numbers
193: greatly reduces the possibility that our results are an artifact of
194: the potential calculation scheme, an issue that has plagued the
195: interpretation of similar results in the past \citep{Zaritsky:88}.
196:
197: \S\,\ref{sec:models} describes the initial models and the $N$-body
198: algorithm. Evolution of the SMBH's orbit is described in detail in
199: \S\,\ref{sec:bhmotion}, and the induced changes in galaxy structure
200: are described in \S\,\ref{sec:profiles}, where the $N$-body models are
201: compared to luminosity profiles of core galaxies. \S\,\ref{sec:times}
202: presents estimates of the SMBH return times in real galaxies, and
203: \S\,\ref{sec:obs} discusses some of the observable consequences of the
204: kicks.
205:
206:
207: \section{Initial models and numerical methods}
208: \label{sec:models}
209:
210: The light profiles of elliptical galaxies and the bulges of spiral
211: galaxies are generally well described in terms of the S{\'e}rsic model
212: \citep{1963BAAA....6...41S, 1968adga.book.....S}, which is a
213: generalization of the \cite{1948AnAp...11..247D,1959HDP....53..275D}
214: law. The most luminous elliptical galaxies depart systematically from
215: the S{\'e}rsic law near the center, where they show evidence for
216: partially depleted stellar cores
217: \citep{Faber:97,Milos:02,Graham:04,ACS:VI}. Formation of a binary
218: SMBH following a galaxy ``major merger'' has been shown to produce
219: cores of roughly the right magnitude \citep{MM:01,Merritt:06},
220: although some observed cores are too large to be easily explained by
221: this model (a point we return to in detail below).
222:
223: As approximate representations of galaxies with binary-depleted cores,
224: we adopt core-S{\'e}rsic models \citep{Graham:03} for our initial conditions.
225: The {\it space} density profile of a galaxy that follows
226: the core-S{\'e}rsic law in projection
227: can be accurately approximated as \citep{2005MNRAS.362..197T}
228: \begin{eqnarray}
229: \rho \left(r\right) & = & \rho^{'} \left[1 + \left(\frac{\rb}{r}\right)^{\alpha}\right]^{\gamma/\alpha}\nonumber \\
230: & & \left[\left(r^{\alpha} + \rb^{\alpha}\right)/\re^{\alpha}\right]^{-p/\alpha}
231: e^{-b \left[\left(r^{\alpha}+\rb^{\alpha}\right)/\re^{\alpha}\right]^{1/n \alpha}}
232: \label{eq:PS1}
233: \end{eqnarray}
234: with
235: \begin{equation}
236: \rho^{'} = \rhob \,2^{\left(p-\gamma\right)/\alpha} \left(\frac{\rb}{\re}\right)^p
237: e^{b \left(2^{1/\alpha} \rb/\re \right)^{1/n}}.
238: \label{eq:PS2}
239: \end{equation}
240: Equation~(\ref{eq:PS1}) is a modification of the Prugniel-Simien model
241: \citep{PS:97}. Here, $\re$ is the effective (half-mass) radius of the
242: projected galaxy; $\rb$ is the break (core) radius; $\rhob$ is the
243: space density at $r=r_b$; and $\alpha$ regulates the sharpness of the
244: transition from core to outer profile. The parameter $n$ describes
245: the curvature of the S{\'e}rsic profile and $b$ and $p$ are fixed
246: functions of $n$ \citep{PS:97,2005MNRAS.362..197T}. Monte-Carlo
247: initial conditions were generated using the scheme of \cite{Szell:05},
248: after including the gravitational potential of a central point
249: particle representing the SMBH.
250:
251: The parameters used for our initial models are listed in
252: Table~\ref{tab:models}. The table also reports names for the
253: different runs based on the adopted ratio of SMBH mass to galaxy mass
254: and initial core radius. Core radii were chosen so as to give initial
255: mass deficits of roughly $\mbh$, as observed for the majority of
256: luminous early-type galaxies \citep{Merritt:06}. We note that
257: $\gamma=0.5$ is the shallowest power-law profile that is consistent
258: with a non-negative, isotropic distribution of stellar velocities
259: around the BH.
260:
261: \begin{table}
262: \begin{center}
263: \caption{Parameters of the initial models.}
264: \label{tab:models}
265: \begin{tabular}{cccccc}
266: \hline
267: name & $n$ & $\alpha$ & $\rb$ & $\gamma$ & $\mbh/\mg$\\
268: \hline
269: A1 & 4.0 & 2.0 & 0.014 & 0.55 & $1.0\times 10^{-3}$ \\
270: A2 & 4.0 & 2.0 & 0.0095 & 0.55 & $1.0\times 10^{-3}$ \\
271: B & 4.0 & 2.0 & 0.027 & 0.55 & $3.0\times 10^{-3}$ \\
272: \hline
273: \end{tabular}
274: \end{center}
275: \end{table}
276:
277: The initial models were evolved using the $\phi$GRAPE numerical
278: integrator \citep{Harfst:07}. This direct-summation code employs a
279: fourth-order Hermite integrator with predictor-corrector scheme and
280: hierarchical time steps. The MPI parallelization strategy is designed
281: to minimize the amount of communication among different computing
282: nodes and to make efficient use of the special-purpose GRAPE hardware.
283: All the simulations presented in this work were performed on the
284: 32-node cluster {\tt
285: gravitySimulator}\footnote{http://wiki.cs.rit.edu/bin/view/GRAPEcluster}
286: at the Rochester Institute of Technology. Most of our simulations
287: used $N=0.5\times 10^6$ equal-mass particles to represent the galaxy
288: although some runs used larger $N$. We set the ratio of BH mass to
289: galaxy mass, $\mbh/M_{\rm gal}$, to be $(1,3)\times 10^{-3}$, typical
290: for observed galaxies \citep{2001MNRAS.320L..30M}. For each model
291: described in Table~\ref{tab:models}, we chose eleven different values
292: of the kick velocity $\vk$ in units of the central escape speed $\ve$:
293: $\vk=(0.1,0.2,..,1.1) \times\ve$; the latter was computed numerically
294: from the initial $N$-body models. In order to guarantee energy
295: conservation, we used a time-step accuracy parameter $\eta = 0.01$.
296: This ensures a relative energy error smaller than one part in $10^6$.
297: An accuracy parameter twice as big would approximately halve the
298: integration time but would result in a relative energy error of
299: $10^{-4}$, which we do not consider acceptable for this study. A
300: softening length $\epsilon = 10^{-4}$ was assigned to both the stars
301: and the BH. Such a small softening length has been shown not to
302: affect even the Brownian motion of a massive particle in models like
303: ours \citep{MBL:07}.
304:
305: Throughout the paper we adopt units according to which the
306: gravitational constant $G$, the effective radius $\re$ in equation
307: (\ref{eq:PS1}), and the total galaxy mass $M_{\rm gal}$ are unity.
308: The models can be scaled to physical units as follows:
309: \begin{mathletters}
310: \begin{eqnarray}
311: \left[T\right] &=& \left(\frac{G\,M_{\rm gal}}{\re^3}\right)^{-1/2} \\
312: &=& 7.75 \times 10^6 {\rm yr}
313: \left(\frac{M_{\rm gal}}{10^{11}\msun}\right)^{-1/2}
314: \left(\frac{\re}{3\,\rm kpc}\right)^{3/2},\\
315: \left[V\right] &=& \left(\frac{G M_{\rm gal}}{\re}\right)^{1/2} \\
316: &=& 378\kms
317: \left(\frac{M_{\rm gal}}{10^{11}\msun}\right)^{1/2}
318: \left(\frac{\re}{3\,\rm kpc}\right)^{-1/2}.
319: \end{eqnarray}
320: \label{eq:units}
321: \end{mathletters}
322:
323:
324: \section{The black hole motion}
325: \label{sec:bhmotion}
326:
327: \subsection {General Remarks}
328: \begin{figure}
329: \begin{center}
330: \includegraphics[width=8.0cm]{f1.eps}
331: \end{center}
332: \caption{BH trajectories in models A1, A2 and B, for $\vk/\ve=0.4$
333: (blue/lower), $0.7$ (green), $0.9$ (red) and $1.1$ (black).}
334: \label{fig:roft}
335: \end{figure}
336: Figure~\ref{fig:roft} (which can be compared with Fig.~1 of
337: \citealp{madauq:04} shows BH trajectories in models A1, A2 and B for
338: $\vk/\ve=(0.4,0.7,0.9,1.1)$. For $\vk \ge \ve$ the black hole escapes
339: the galaxy on an unbound orbit.
340: \begin{figure}
341: \begin{center}
342: \includegraphics[width=8.5cm]{f2.eps}
343: \end{center}
344: \caption{Maximum displacement of the BH from the galaxy center. The
345: data points show the results from the simulations while the lines
346: are estimates in the absence of dynamical friction. The dashed
347: lines represent numerical estimates from the computation of the
348: potential of the $N$-body system at time $t = 0$ while the dotted
349: lines represent theoretical estimates from the analytic expression
350: of the potential in a core-S{\'e}rsic model.}
351: \label{fig:rmax}
352: \end{figure}
353: The maximum displacement of the BH ($r_{\rm max}$) is shown in
354: Figure~\ref{fig:rmax}. The data from the simulations (points) are
355: compared to theoretical (dotted lines) and numerical (dashed lines)
356: estimates of $r_{\rm max}$ in the absence of dynamical friction. The
357: theoretical and numerical estimates are obtained from the initial
358: $N$-body data by assuming conservation of total energy for the BH:
359: $r_{\rm max}$ is the distance at which the gravitational potential of
360: the system equals the initial total energy of the BH. For the
361: theoretical solution we use the expression of the potential in a
362: core-S{\'e}rsic model \citep[see equations 7 through 13
363: of][]{2005MNRAS.362..197T} while for the numerical solution we compute
364: the potential at different radii from the $N$-body data.
365: The two estimates are for practical purposes indistinguishable.
366:
367: Dynamical friction affects the maximum displacement of the BH only for
368: moderately large kicks, where the data points appear systematically
369: lower than the theoretical curves. Values of $r_{\rm max}$ larger than
370: the expected turning points in the first orbit are due to the
371: rapidly-expanding core.
372:
373: During the initial outward journey, dynamical friction does not
374: strongly influence the motion of the BH, and the maximum displacement
375: is similar to that of an energy-conserving orbit. We note that a kick
376: velocity larger than about $0.3\,\ve$ is necessary to bring the BH
377: beyond the core.
378: \begin{figure}
379: \begin{center}
380: \includegraphics[width=8.5cm]{f3.eps}
381: \end{center}
382: \caption{{\it Upper panel:} Specific energy of the BH particle
383: versus time in Model A1 with $\vk=0.9\ve$.
384: Almost all of the energy loss occurs during passages through the core.
385: {\it Lower panel:} Mean density in a sphere of radius 0.05 centered on
386: the point of maximum density in the core of the galaxy
387: (excluding the BH).}
388: \label{fig:eoft}
389: \end{figure}
390: Due to the combined effect of the kick and dynamical friction, the BH
391: displays a damped oscillatory motion. The number of radial
392: oscillations increases with $\vk$; for $\vk=0.9\,\ve$ the BH
393: experiences $\sim 5$ full radial oscillations before returning to the
394: core. Almost all of the energy loss to dynamical friction takes place
395: during the short intervals that the BH passes through the core. This
396: is shown in Figure~\ref{fig:eoft} which plots the evolution of the BH
397: specific energy $E$ in Model A1 with $\vk=0.9\ve$, where
398: \begin{equation} E \equiv \frac{V^2}{2} - \sum_{i=2}^N
399: \frac{m_i}{\sqrt{(\mathbf{x}_i-\mathbf{X})^2 + \epsilon^2}}
400: \label{eq:defe}
401: \end{equation}
402: and the summation is over the ``star'' particles\footnote{Unless
403: otherwise noted, upper-case variables $X$ and $V$ refer to the BH
404: particle while lower-case symbols are reserved for the star
405: particles.}. The energy lost during the initial emergence from the
406: core appears to be less than during subsequent passages, suggesting
407: that dynamical friction requires a finite time to ``turn on'' after
408: the kick. During the first few oscillations, the BH's motion remains
409: essentially rectilinear, but eventually the $Y$- and $Z$-components of
410: the motion become important due to non-sphericities in the galaxy
411: potential and also to perturbations from stars. At late times, the
412: BH's motion is essentially random, similar to that of a Brownian
413: particle in a fluid. Figure~\ref{fig:eoft} also shows the mean
414: density in a sphere of fixed radius whose center is located at the
415: estimated density peak (computed via the algorithm described in
416: \S\,\ref{sec:phaseI}). The core density decreases rapidly following
417: the initial ejection, then more gradually as the BH returns again and
418: again to the core, losing energy to the stars each time.
419:
420: \begin{figure}
421: \begin{center}
422: \includegraphics[width=8.5cm]{f4.eps}
423: \end{center}
424: \caption{Stellar mass bound to the BH in the initial models.}
425: \label{fig:bound}
426: \end{figure}
427: Figure~\ref{fig:bound} shows the mass in stars bound to the BH at
428: $t=0$. The bound mass was computed by counting all the stars, within
429: the influence radius $r_h$, which formed a bound two-body system with
430: the BH particle. The influence radius was defined as the radius
431: containing a mass in stars equal to twice $\mbh$. The bound mass
432: decreases steeply with $\vk$, as noted in earlier studies
433: \citep{MMFHH:04,Boylan:04}, and is ignorable for $\vk\gtrsim 0.6\ve$.
434:
435: In all cases where the kick velocity was large enough to remove the BH
436: completely from the core (i.e. $\vk\gtrsim 0.3 \ve$), we observed
437: three distinct regimes of the motion. In Phase I, the BH's motion is
438: well predicted by Chandrasekhar's dynamical friction theory, after
439: taking into account the changing size of the galaxy core where most of
440: the friction occurs. This is the phase illustrated in
441: Figure~\ref{fig:roft}; in Figure~\ref{fig:eoft}, Phase I extends until
442: $t\approx 20$. Phase II begins roughly when the amplitude of the BH's
443: motion had decayed to the size of the core. In this phase, the energy
444: of the BH's orbit continues to decay but with a much longer time
445: constant than predicted by Chandrasekhar's formula. The BH and the
446: core oscillate about their common center of mass in this regime. In
447: Phase III, the BH's energy has dropped to the thermal level. Phase II
448: is generally longer than Phase I, and this would presumably be even
449: more true in real galaxies since the amplitude of thermal oscillations
450: is much lower than in our simulations implying a longer time to reach
451: the Brownian regime. We discuss these three regimes in detail below.
452:
453: \subsection {Phase I}
454: \label{sec:phaseI}
455:
456: The extent of Phase I is clearly indicated in the plots of BH energy
457: vs. time (e.g. Fig.~\ref{fig:eoft}): a distinct ``knee'' appears in
458: the $E(t)$ curves marking the end of this phase. Values of $T_I$, the
459: elapsed time from the kick until the end of Phase I, are given in
460: Table~\ref{tab:times}.
461:
462: We compared the evolution of the BH's motion in Phase I with the
463: predictions of Chandrasekhar's dynamical friction theory
464: \citep{1943ApJ....97..255C}. Such comparisons are problematic since
465: much of the energy exchange between BH and stars occurs during
466: passages through the galaxy's core, and the core density changes
467: significantly with time due to the BH's motion. We dealt with this
468: problem by breaking the BH's motion into segments, each containing one
469: passage through the center, and assuming that the galaxy's density
470: remained constant during each segment.
471:
472: Chandrasekhar (1943) derived his expression for the dynamical friction
473: acceleration $F_{\rm df}$ assuming an infinite, homogeneous and
474: unchanging background of perturbers (stars).
475: In the limit that the mass of
476: the heavy object greatly exceeds the masses of the stars,
477: the acceleration is predicted to be
478: \begin{equation} F_{\rm df} \approx -2 \pi G^2 \rho\,
479: \mbh \, \ln(1+\Lambda^2)\, \vbh^{-2} N(<\vbh,r),
480: \label{eq:df}
481: \end{equation}
482: where $\rho(\mathbf{r})$ is the mass density of stars at the BH's
483: position,
484: $(1/2)\ln(1+\Lambda^2)$ is the Coulomb logarithm,
485: $\vbh$ is the BH's instantaneous velocity,
486: and $N(<\vbh,{\mathbf r})$ is the fraction of stars at
487: $\mathbf{r}$ that are moving (in the frame of the galaxy) with
488: velocities less than $\vbh$.
489:
490: Some care must be taken in the definition of the Coulomb logarithm.
491: One commonly writes
492: \begin{equation}
493: \ln\left(1+\Lambda^2\right) \approx 2\ln\Lambda \approx 2\ln (p_{\rm max}/p_{\rm min})
494: \end{equation}
495: where $p_{\rm min}$ and $p_{\rm max}$ are the minimum and maximum
496: effective impact parameters of the stars that contribute to the
497: frictional force, and $p_{\rm max}\gg p_{\rm min}$.
498: However, $p_{\rm min}$ depends on the field-star velocity
499: \citep{White:49,Merritt:01}
500: and $p_{\rm max}$ is likewise ill-defined
501: since a realistic stellar system is inhomogeneous and
502: has no outer boundary.
503:
504: Numerous $N$-body simulations have been carried out to evaluate
505: Chandrasekhar's formula in the case of a massive particle inspiraling
506: toward the center of a galaxy
507: \citep{White:83,Bontekoe:87,Bontekoe:88,Weinberg:89,Cora:97,Bertin:03}.
508: Early work was typically based on approximate $N$-body schemes and the
509: results were often discrepant from study to study \citep{Zaritsky:88}.
510: These differences appear to have been resolved in the last few years
511: through the use of direct-summation codes
512: \citep{Spinnato:03,Merritt:06}, which consistently find
513: $4\lesssim\ln\Lambda\lesssim 6$ for inspiral of massive point
514: particles, on circular or near-circular orbits, into the centers of
515: galaxies with steeply-rising density profiles.
516: Fewer experiments have been done with highly eccentric orbits,
517: although \cite{Just:05}, using an approximate method,
518: find $2\lesssim\ln\Lambda\lesssim 3$ for orbits with moderate eccentricities.
519:
520: In general, we expect the effective value of $\ln\Lambda$ to be smaller
521: for radial orbits than for circular motion.
522: The dynamical friction force arises from a polarization of the stellar
523: density which produces an over-dense region, or wake, behind the
524: massive object \citep{Mulder:83}. A finite time, of order a galaxy
525: crossing time, is presumably required for this wake to be set up. In
526: the case of a gradually-decaying circular orbit, the galaxy is able to
527: reach a quasi-steady state after a few orbits of the massive object.
528: In our case, the position and velocity of the BH are changing
529: dramatically over one crossing time, so that the wake never has a
530: chance to establish its steady-state amplitude; indeed just after
531: apocenter passages, the over-dense region can be seen to lie in {\it
532: front} of the BH.
533:
534: In order to determine the effective value of $\ln\Lambda$ in the
535: $N$-body integrations, we computed BH trajectories using
536: Chandrasekhar's formula (equation\,\ref{eq:df}) with various values of
537: the $\ln\left(1+\Lambda^2\right)$ term (henceforth written simply as
538: $2\ln\Lambda$) and compared them with the $N$-body trajectories. The
539: following procedure was followed.
540:
541: 1. The density center of the galaxy moves slightly with respect to the
542: origin of the coordinates due to transfer of momentum from the kicked
543: BH to the galaxy. In order to accurately determine the distance of the
544: BH from the galaxy center as a function of time, we recorded full
545: snapshots of the particle positions at frequent intervals, then used
546: the Casertano-Hut (1985) algorithm to find the density center of the
547: stars in each snapshot. A smoothing spline was fit through the
548: measured positions to give a continuous estimate of the center
549: displacement as a function of time, and this displacement was
550: subtracted from the BH positions. (The instantaneous velocity of the
551: density center was ignored, which is a good approximation at least
552: until the end of Phase I.) The resulting correction was at most $\sim
553: 0.02$; at late times the displacement reached a constant value since
554: the center-of-mass velocity of the system was zero by construction.
555:
556: 2. In order to apply Chandrasekhar's formula we needed to specify the
557: galaxy model. The galaxy's mass distribution changes with time due to
558: the BH's motion; most of this change takes place in the core just
559: after the BH passes through. We therefore fixed all the parameters in
560: equation~(\ref{eq:PS1}) except for the core radius $r_b$. We
561: determined the effective value of $r_b$ at the discrete times when the
562: BH passed through the galaxy center by assuming a flat core
563: ($\gamma=0$) and finding the value of $r_b$ such that the mass
564: contained within $r_b$ according to equation~(\ref{eq:PS1}), with
565: $\alpha=2$, was the same as the mass in the $N$-body model in a sphere
566: of radius $r_b$ centered on the BH. This procedure was always found
567: to yield a unique $r_b$ and accurately recovered the known value of
568: $r_b$ in the initial models.
569:
570: 3. BH trajectories were then computed in a piecewise fashion using
571: Chandrasekhar's formula, starting from one extremum in the BH
572: displacement and continuing until the next extremum, using the value
573: of $r_b$ corresponding to the central passage lying between the two
574: extrema. This was repeated for several values of $\ln\Lambda$. We
575: used equation~(5) of \cite{Szell:05} to compute $N(<\vbh,r)$ in
576: equation~(\ref{eq:df}) from the assumed $\rho(r)$.
577:
578: \begin{figure}
579: \begin{center}
580: \includegraphics[width=8.5cm]{f5.eps}
581: \end{center}
582: \caption{Comparison between BH trajectories computed via the $N$-body
583: integrations (open circles) and via Chandrasekhar's formula
584: (\ref{eq:df}) (lines). The $N$-body models were A1 ($\mbh=0.001$) with
585: $\vk=0.7\,\ve$ (a) and B ($\mbh=0.003$) with $\vk=0.8\,\ve$ (b).
586: Theoretical trajectories were computed in a piecewise manner,
587: starting from extrema in the BH's trajectory (vertical solid lines)
588: and continuing until
589: the next extremum; the core radius $r_b$ of the galaxy model was
590: adjusted as described in the text to give the same core density as
591: in the $N$-body model at the time when the BH passed through the
592: center. Horizontal dashed lines show the adopted values of $r_b$. Line
593: colors/styles correspond to different values of $\ln\Lambda$: $1$
594: (blue/solid), $2$ (magenta/dashed), $3$ (red/dash-dotted), $4$
595: (black/dotted).}
596: \label{fig:comp}
597: \end{figure}
598:
599: Figure~\ref{fig:comp} shows the results for Model A1 with
600: $\vk/\ve=0.7$ and Model B with $\vk/\ve=0.8$. During each inward leg
601: of the trajectory, the dynamical friction force hardly affects the
602: motion; only when passing through the dense center is the motion
603: significantly non-ballistic. (This could be seen already in
604: Figures\,\ref{fig:rmax} and~\ref{fig:eoft}.) The best-fit value of
605: $\ln\Lambda$ was found to lie in the range
606: $2\lesssim\ln\Lambda\lesssim 3$, and for such values, Chandrasekhar's
607: formula did a good job of reproducing the motion. We found no
608: evidence of a systematic change in the effective value of $\ln\Lambda$
609: from one time interval to the next.
610:
611: \subsection{Phase III}
612: \label{sec:phaseIII}
613:
614: The BH trajectories in Figure~\ref{fig:comp} are displayed until the
615: amplitude of the oscillations has decayed down to roughly the core
616: radius. As discussed above, the BH's motion is well predicted by
617: Chandrasekhar's dynamical friction formula in this regime. Shortly
618: after returning to the core, however, the BH's motion was found to
619: depart strikingly from the predictions of Chandrasekhar's formula. A
620: detailed discussion of the motion in ``Phase II'' is presented below.
621: Before doing so, we consider the motion of the BH at still later
622: times, ``Phase III,'' when it has reached thermal equilibrium with the
623: stars.
624:
625: \begin{figure}
626: \begin{center}
627: \includegraphics[width=8.5cm]{f6.eps}
628: \end{center}
629: \caption{Squared BH velocity in seven $N$-body integrations of Model
630: B. For $\vk\gtrsim 0.4\ve$ the BH moves completely out of the
631: core before falling back. Ticked, horizontal lines demarcate
632: Phase II. Blue (dashed) lines show $\langle V^2\rangle$ during
633: Phase III, and red (dotted) lines show the mean square velocity
634: predicted by equation~(\ref{eq:brown}), which assumes that the BH
635: particle has reached thermal equilibrium with the stars in its
636: vicinity. }
637: \label{fig:voft}
638: \end{figure}
639:
640: Figure~\ref{fig:voft} shows the squared velocity of the BH,
641: $V^2=V_x^2+V_y^2+V_z^2$, over the full integration interval, for kick
642: velocities $\vk\ge 0.3\ve$ in Model B. For $\vk\gtrsim 0.4\ve$ the BH
643: moves substantially beyond the core during its first oscillation
644: (Fig.~\ref{fig:rmax}). At late times, the motion of the BH in each of
645: these integrations appears to be stochastic (i.e. non-quasi-periodic)
646: but with roughly constant amplitude.
647:
648: The dashed (blue) lines in this figure show $\left\langle
649: V^2\right\rangle$, the mean square velocity of the BH averaged over
650: Phase III. (The precise definition of the start of Phase III is given
651: below.) Also shown (dotted red lines) are estimates of the expected
652: value of $\left\langle V^2\right\rangle$ for the BH once it reaches
653: statistical equilibrium with the stars. The latter velocity, $V_{\rm
654: Brown}^2$, was computed using
655: \begin{equation} V_{\rm Brown}^2 = 3 \frac{m_\star}{\mbh} {\tilde\sigma}^2.
656: \label{eq:brown}
657: \end{equation}
658: Equation~(\ref{eq:brown}) equates the kinetic energy of the BH
659: with the mean kinetic energy of a single star in the core. The
660: quantity $\tilde\sigma$ is defined as the 1D velocity dispersion of
661: stars within a sphere of radius $K\times r_h$ centered on the BH, with
662: $r_h$ the BH's influence radius (the radius containing a mass in stars
663: equal to twice $\mbh$) and $K$ a constant of order unity.
664: \cite{MBL:07} used $N$-body simulations to evaluate $K$ for massive
665: particles at the centers of galaxies with power-law nuclear density
666: profiles, $\rho\sim r^{-\gamma}$. They found that $K$ increases
667: slowly with decreasing $\gamma$, to $K\approx 0.8$ when $\gamma=0.5$.
668: We set $K=1$ when computing $V_{\rm Brown}$ in Figure~\ref{fig:voft};
669: the agreement with the measured values is quite good, confirming that
670: the BH behaves as a Brownian particle in Phase III.
671:
672: \begin{figure}
673: \begin{center}
674: \includegraphics[width=8.5cm]{f7.eps}
675: \end{center}
676: \caption{RMS amplitude of the BH oscillations in the Brownian
677: regime, Phase III, for models A1 (black/circles), A2
678: (blue/squares) and B (red/triangles).}
679: \label{fig:rbrown}
680: \end{figure}
681:
682: Figure~\ref{fig:rbrown} shows the rms amplitude of the BH's motion
683: averaged over Phase III.
684: Since the density center of the galaxy drifts, as described above,
685: smoothing splines were first fit to the ${\bf X}(t)$ values
686: for the BH and the rms deviations were computed with respect
687: to the smoothed trajectories.
688: Figure~\ref{fig:rbrown} shows a general trend of increasing $R_{\rm rms}$ with
689: decreasing core density, as expected if the motion in this
690: regime obeys the virial theorem,
691: \begin{equation}
692: \langle V^2 \rangle \approx \frac{4}{3} \pi G \rho_c \langle R^2\rangle.
693: \label{eq:rbrown}
694: \end{equation}
695: This relation \citep[cf.][]{BW:76} assumes a constant-density
696: core, ignores the back-reaction of the BH's motion on the stars, and
697: ignores any coupling between random gravitational perturbations from
698: the stars and the quasi-periodic motion of the BH in the smooth
699: potential of the core. Nevertheless, equation~(\ref{eq:rbrown}) was
700: found to reproduce the measured $R_{\rm rms}$ values in
701: Figure~\ref{fig:rbrown} quite well if $\rho_c$ was defined as the mean
702: density of stars within $r_h$. Fluctuations in $R_{\rm rms}$ about
703: the mean relation in Figure~\ref{fig:rbrown} appear to be due
704: primarily to fluctuations in $V_{\rm rms}$ and would presumably be
705: smaller if the $R_{\rm rms}$ values were averaged over longer time
706: intervals. The near agreement between the $R_{\rm rms}$ values for
707: the runs with small and large $\mbh$ is a consequence of the larger
708: core size / lower core density in runs with larger $\mbh$, which
709: compensates for the lower $\langle V^2\rangle\propto \mbh^{-1}$.
710:
711: We note here that the amplitude of the BH's Brownian motion is always
712: a factor 10 or more smaller than the final core radii of the models
713: (Table~\ref{tab:fits}). This implies that the motion of the BH when
714: it first returns to the core -- at the start of Phase II -- should not
715: be appreciably affected by discreteness effects, i.e. by perturbations
716: from individual stars. This conclusion is confirmed below.
717:
718: We note also that the amplitude of Brownian oscillations of BHs in
719: real galaxies (expressed as a fraction of the galaxy effective radius,
720: say) would be smaller than in our models by the factor $\sim
721: \sqrt{(M_{\rm gal}/m_\star)/N}$, i.e. $\sim 50$ for $M_{\rm
722: gal}=10^9M_\odot$ and $\sim 500$ for $M_{\rm gal}=10^{11}M_\odot$.
723: The time required for a BH to reach these lower kinetic energies would
724: also presumably be longer than in our simulations, as discussed in
725: more detail below.
726:
727: \subsection{Phase II}
728: \label{sec:phaseII}
729:
730: As noted above, the motion of the BH after returning to the core, and
731: before reaching the Brownian regime, is not well described by
732: Chandrasekhar's formula. Here we consider the motion in this regime
733: (``Phase II'').
734:
735: Figure~\ref{fig:voft} reveals the following qualitative features.
736:
737: 1. The motion in Phase II is essentially oscillatory, with a period
738: similar to that at the end of Phase I, i.e. roughly equal to the
739: period of oscillation of a test particle moving in the stellar core.
740:
741: 2. There is evidence of additional frequencies affecting the BH's
742: motion. For instance, the amplitude of the oscillations sometimes
743: appears to {\it increase} temporarily over several periods in a manner
744: suggestive of beats.
745:
746: 3. Averaged over many periods, the mean amplitude of the oscillations
747: decays, but with a time constant that is much longer than observed
748: toward the end of Phase I.
749:
750: 4. Near the end of Phase II, the motion becomes increasingly
751: stochastic, presumably due to perturbations from individual stars.
752: Eventually the BH rms velocity falls to the Brownian (thermal) level
753: marking the start of Phase III.
754:
755: 5. Phase II always begins roughly when the stellar mass interior to
756: the BH's orbit is equal to $\mbh$. When $\vk\lesssim 0.3\ve$, the BH
757: never escapes the core, and its motion appears to transition directly
758: from Phase I to Phase III.
759:
760: Based on Figure~\ref{fig:voft}, the elapsed time in Phase II can be
761: substantially longer than the time spent in Phase I. Understanding
762: the character of the motion in this regime is therefore crucial for
763: predicting the expected displacement of a supermassive BH in a real
764: galaxy following a kick.
765:
766: We begin by considering a simple model for damped oscillations of a
767: massive particle in a constant-density core. While this model will
768: fail to quantitatively reproduce the motion in Phase II, it provides a
769: useful framework for discussing what is observed in the simulations.
770:
771: In the absence of dynamical friction, and neglecting the influence
772: of the massive particle's presence on core structure,
773: the motion of the massive particle is simple harmonic oscillation
774: with frequency
775: $\omega_c = \sqrt{(4\pi/3)G \rho_c}$;
776: $\rho_c$ is the core density, assumed constant within
777: a radius $r_c$.
778: To this motion we add the acceleration due to dynamical friction.
779: If the velocity distribution of the stars that produce the friction
780: is Maxwellian with 1D velocity dispersion $\sigma_c$, and if
781: the BH's velocity satisfies $V\ll\sigma_c$,
782: the resulting equation of motion in any coordinate $x_i$ is
783: \begin{equation} \ddot{X_i} + \tdf^{-1}\dot{X_i} + \omega_c^2 X_i = 0
784: \label{eq:df1}
785: \end{equation}
786: where
787: \begin{equation}
788: \tdf = \frac{3}{8} \sqrt{\frac{2}{\pi}} \frac{\sigma_c^3}{ G^2\rho_c\mbh\ln\Lambda}
789: \end{equation}
790: is the dynamical friction damping time \citep{Merritt:85}.
791: The condition for underdamped oscillations is $2\omega_c\tdf> 1$,
792: where
793: \begin{mathletters}
794: \begin{eqnarray}
795: 2\omega_c\tdf &=& \frac{\sqrt{6}}{2}
796: \frac{\sigma_c^3}{G^{3/2}\rho_c^{1/2}\mbh\ln\Lambda}
797: \label{eq:damp} \\
798: &=& \frac{\sqrt{6\pi}}{9} \form^3
799: \frac{M_c}{\mbh\ln\Lambda},
800: \end{eqnarray}
801: \end{mathletters}
802: with $M_c \equiv (4/3) \pi \rho_cr_c^3$ the core mass;
803: the second relation uses the ``core-fitting'' formula
804: of \citet{RPKK:72},
805: \begin{equation}
806: \sigma_c^2 = \form^2 \frac{4\pi}{9}G\rho_c r_c^2.
807: \label{eq:King}
808: \end{equation}
809: $\form \approx 2$ for our models.
810: Thus
811: \begin{equation}
812: 2\omega_c \tdf \approx 4 \frac{M_c}{\mbh\ln\Lambda}
813: \end{equation}
814:
815: \begin{figure}
816: \vspace*{-1mm}
817: \begin{center}
818: \includegraphics[width=8.5cm]{f8.eps}
819: \end{center}
820: \caption{Evolution of the BH kinetic energy in a series of integrations
821: of model B with various $N$, and $\vk=0.6\ve$.
822: {\it Top panel:} Squared velocity of the BH versus time.
823: Dashed lines at the right show the predicted values of $V^2$
824: in the Brownian regime (eq.~\ref{eq:brown}).
825: {\it Bottom panel:} Binned values of $V^2$ in Phases II and III.
826: Dotted lines are least-squares fits to the binned data.
827: These fits are plotted until the time at which they
828: intersect the Brownian $V^2$; these times are marked
829: by the vertical solid lines.
830: The latter are found to be spaced with roughly constant
831: separation indicating that the time required for the BH to reach
832: thermal equilibrium with the stars increases roughly as $\ln N$.}
833: \label{fig:tests}
834: \end{figure}
835:
836: In our simulations (and in real galaxies), the right hand side of this
837: expression is $\gtrsim 1$, since core masses are $\sim $ a few $\mbh$
838: \citep{Merritt:06} and $2\lesssim \ln\Lambda\lesssim 3$
839: (\S\,\ref{sec:phaseI}). It follows that the motion of the BH should
840: be under-damped, though not far from critically damped, after it
841: re-enters the core. The solutions to equation (\ref{eq:df1}) in the
842: under-damped regime are
843: \begin{equation}
844: X_i(t) = A_i\,e^{-t/2\tdf} \sin\left(\omega_c t + \phi_i\right).
845: \label{eq:xioft}
846: \end{equation}
847: Writing $\Theta\equiv 2\omega_cT_{\rm df}\gtrsim 1$ and
848: $T_c\equiv 2\pi/\omega_c$, the energy decay time is predicted to be
849: $T_{\rm df}=(\Theta/4\pi)T_c$, i.e. shorter than the orbital period.
850: Such short decay times are in fact observed near the end of Phase I
851: (Figure~\ref{fig:comp}).
852:
853: However, Figure~\ref{fig:voft} shows that this is not the case in
854: Phase II: the mean damping time is substantially longer than an
855: orbital period. The abrupt decrease in the energy dissipation rate at
856: the start of Phase II can also be seen in Figure~\ref{fig:eoft}(a).
857:
858: A possible explanation for the slower damping in Phase II is
859: discreteness effects: perturbations from individual stars, some of
860: which act to accelerate the BH, become increasingly competitive with
861: mean-field effects (including dynamical friction) as the BH moves more
862: slowly. Indeed, in the Brownian regime (Phase III), the accelerating
863: perturbations are equally as strong, in a time-averaged sense, as
864: dynamical friction. While the amplitude of the BH oscillations at the
865: onset of Phase II is always much greater than the Brownian amplitude
866: in these simulations (cf.~Fig.~\ref{fig:rbrown} and the accompanying
867: discussion), it is still conceivable that discreteness effects are
868: responsible for the anomalously slow decay of the BH's orbit at this
869: time.
870:
871: To securely rule out this possibility, we repeated the integration of
872: model B with $\vk=0.6\ve$, increasing $N$ up to $N=2\times 10^6$.
873: Figure~\ref{fig:tests} shows the results. The slowly-damped
874: oscillations in Phase II are clearly not an artifact of a too-small
875: $N$. In all cases, for instance, the fifth extremum in $V^2$ (which
876: occurs at $t\approx 1.42$) is comparable or greater in amplitude to
877: the fourth extremum (at $t\approx 1.16$), rather than being much lower
878: in amplitude as would be expected from the above analysis or from
879: Figure~\ref{fig:comp}. We also carried out a number of tests varying
880: the integration time-step parameter $\eta$; again, no systematic
881: dependence of the evolution in Phase II on this parameter was
882: observed.
883:
884: Particularly striking in Figure~\ref{fig:tests} is the accurately
885: exponential decay of the BH's kinetic energy throughout Phase II; this
886: is clearest in the simulation with largest $N$, where the exponential
887: damping continues over two decades in energy. We note again that an
888: exponentially decaying energy is predicted by the simple model just
889: presented, but the model predicts a much shorter time constant than
890: what is observed in the $N$-body simulations.
891:
892: \begin{figure}
893: \begin{center}
894: \includegraphics[width=8.5cm]{f9.eps}
895: \end{center}
896: \caption{Core-BH oscillations in Phase II. This is the $N=2\times
897: 10^6$ integration of Model B shown as the filled (red) circles in
898: Fig.~\ref{fig:tests}. Contours are separated by $0.034$ in
899: $\log_{10}$ of the projected density. Filled circles mark the BH
900: and crosses mark the approximate location of the (projected)
901: stellar density maximum. Times are
902: $t=2.1875,2.21875,2.25,...2.46875$, increasing from upper left to
903: lower right. The elapsed time in this figure spans approximately
904: 1/2 oscillation period of the BH. }
905: \label{fig:cont}
906: \end{figure}
907:
908: Figure~\ref{fig:cont} suggests why Chandrasekhar's (1943) formula
909: might break down in Phase II. The approximation of a stationary
910: galaxy is strongly violated in this regime. The galaxy's density
911: center oscillates with opposite phase to the BH, and with roughly the
912: same frequency and amplitude. This is consistent with the observation
913: that Phase II always begins roughly when the mass in stars inside the
914: BH's orbit is similar to $\mbh$. Evidently, in this regime, the BH
915: and the core oscillate about their common center of mass as a
916: two-body system. Chandrasekhar's derivation, which assumed a body on
917: a linear trajectory through an infinite homogeneous medium, is
918: unlikely to apply to oscillations like those in Figure~\ref{fig:cont},
919: since the BH is periodically accelerated, then decelerated, by the
920: density peak. The rate at which such oscillations decay is known to
921: be sensitively dependent on resonant interactions \citep{TW:84} and
922: can be arbitrarily low \citep{Louis:88,Sridhar:89,SN:89,MFR:90},
923: although we are not aware of any theoretical treatment that is
924: directly applicable to oscillations like those in
925: Figure~\ref{fig:cont}.
926:
927: Ours is not the first $N$-body study to observe persistent
928: oscillations of massive objects at the centers of $N$-body models.
929: \cite{Miller:92} and \cite{Miller:96} reported a series of $N$-body
930: integrations, using a grid-based code, of a disk at the center of an
931: axisymmetric galaxy model. They observed what appeared to be
932: over-stable oscillations of particles initially at rest near the
933: center of the disk; the oscillation frequency was roughly
934: $\sqrt{(4\pi/3)G\rho_c}$ and the maximum amplitude was roughly the
935: size of the core. All of these features are characteristic of the
936: oscillations that we observe in Phase II. \cite{Miller:92} also
937: reported ``a couple of experiments in which a massive object was put
938: into orbit within a galaxy model,'' presumably near the center, and
939: observed ``residual oscillations'' with amplitude roughly equal to the
940: radius at which the enclosed mass was equal to the object's mass,
941: again similar to what we observe. \cite{Miller:92} briefly describe a
942: model for the oscillations, in which periodic motion of the core as a
943: whole, at roughly the same frequency as core internal frequencies,
944: drives the oscillations.
945:
946: A number of other $N$-body studies have noted a decrease in the
947: effective value of $\ln\Lambda$ once a massive object has spiraled
948: into a constant-density core. Typically the observed decrease is
949: modest, a factor$\sim 2-3$ or so
950: \citep{Bontekoe:87,Bontekoe:88,Weinberg:89,Cora:97}, although one
951: recent study \citep{Read:06} found a nearly complete disappearance of
952: dynamical friction after the infalling particle reached the core. Read
953: et al. proposed that the apparent vanishing of the dynamical friction
954: force in their simulations could be explained by the degeneracy of
955: orbital frequencies in the harmonic-oscillator potential corresponding
956: to a precisely flat, central density profile. In such models, Read et
957: al. found that the disappearance of the dynamical friction force was
958: critically dependent on whether the plane defined by the inspiralling
959: particle's orbit remained fixed; precession, induced e.g. by
960: finite-$N$ perturbations, caused dynamical friction to turn on again,
961: though at a rate much slower than expected from Chandrasekhar
962: friction. While Read et al. did consider the effect of varying the
963: initial log-slope of the background distribution, their models were
964: always spherical. The cores in our models are not precisely flat nor
965: are our models precisely spherical (once the BH particle has been
966: ejected) and these differences (coupled with the fact that the
967: gravitational potential of the core is highly oscillatory in Phase II)
968: may explain why we do not observe the dramatic stalling reported by
969: Read et al. In any case, the apparent lack of an $N$-dependence in
970: our simulations (Figure~\ref{fig:tests}) suggests that the critical
971: difference between our results and those of Read et al. is not
972: particle number.
973:
974: The Phase II oscillations were clearly visible in every integration
975: with $\vk\ge 0.4\ve$. For $\vk=0.3\ve$ there were hints of a delayed
976: return to the Brownian regime in some of the integrations
977: (e.g. Fig.~\ref{fig:voft}) but not to the extent that we were able to
978: estimate damping times. We could not detect the Phase II oscillations at all
979: for $\vk\le 0.2\ve$; in these integrations, the BH kinetic energy
980: appears to drop very rapidly after the return to the core, more or
981: less as expected based on the analytic model presented above or by an
982: extrapolation of the behavior in Phase I. In any case, we assume in
983: the remainder of this paper that the Phase II oscillations are absent
984: when $\vk\le 0.3\ve$. Integrations with much larger $N$ might modify
985: this conclusion.
986:
987: The occasional {\it increase} in the amplitude of the Phase II
988: oscillations, which is seen in virtually all the integrations, is
989: suggestive of a dynamical instability \citep{Tremaine:05}. However an
990: instability would presumably act even in the case of small kicks,
991: while as noted above, Phase II oscillations appear to be absent for
992: $\vk\lesssim 0.3\ve$. We speculate that the BH must be kicked
993: completely out of the core in order for the BH-core oscillations to be
994: excited, as suggested by \cite{Miller:92}. The roughly sinusoidal
995: variations in the envelope of $V^2(t)$, with a much lower frequency
996: than $\omega_c$, could naturally be explained in terms of beating,
997: e.g. between the frequency of motion in the core and the frequency at
998: which the core itself oscillates in the galactic potential.
999:
1000: Figures~\ref{fig:voft} and~\ref{fig:tests} suggest that the core-BH
1001: oscillations in Phase II decay roughly as an exponential in time, at
1002: least when viewed through a window of several orbital periods or
1003: longer. We investigated a number of ways to quantify the time
1004: constant $\tau$ associated with the energy damping:
1005:
1006: 1. Plots of BH energy versus time (equation~\ref{eq:defe},
1007: Fig.~\ref{fig:eoft}) were found not to be very useful in this regard
1008: since the total energy is dominated by the potential energy which
1009: exhibits fairly large fluctuations from time step to time step.
1010:
1011: 2. In a constant-density core, the unperturbed motion is simple
1012: harmonic oscillation with frequency $\omega$ and energy
1013: $$
1014: E_{\rm SHO} = \frac{1}{2} \sum_{i=1}^3 \left(\omega^2 X_i^2 + V_i^2\right).
1015: $$
1016: We determined the dominant frequency of the BH's motion in Phase II
1017: by carrying out discrete Fourier transforms of the complex functions
1018: $X_i(t)+iV_i(t)$ and constructing power spectra \citep[e.g.][]{Laskar:90}.
1019: Least-squares fits to $\ln E_{\rm SHO}$ vs. $t$ were then carried out
1020: to find the damping time constant. This approach was reasonably
1021: objective and robust, but can be criticized on the grounds that the
1022: core density is not constant and the density center is moving with time
1023: (Fig.~\ref{fig:cont}), making the interpretation of $E_{\rm SHO}$
1024: problematic.
1025:
1026: \begin{figure}
1027: \begin{center}
1028: \includegraphics[width=8.5cm]{f10.eps}
1029: \end{center}
1030: \caption{Energy-decay time constants $\tau$ for the BH in Phase II,
1031: for models A1 (black/circles), A2 (blue/squares) and B (red/triangles).
1032: }
1033: \label{fig:tauomega}
1034: \end{figure}
1035:
1036: 3. Given the difficulties with evaluating and interpreting the total
1037: energy of the BH, we chose in the end to quantify the energy damping
1038: purely in terms of the the BH's kinetic energy. As noted above
1039: (Fig.~\ref{fig:tests} and associated text), $V^2(t)$ exhibits a nicely
1040: exponential decay with a well-defined time constant, and the decay is
1041: observed to continue over $\sim 2$ decades in kinetic energy in the
1042: case of the simulation with the largest $N$, until the BH's kinetic
1043: energy reaches the Brownian value. We evaluated the associated time
1044: constant by carrying out least-squares fits of $\ln V^2$ to time,
1045: yielding the coefficients $(V_I^2, \tau)$ in the expression
1046: \begin{equation}
1047: V^2(t) \approx V_I^2\,e^{-\left(t-T_I\right)/\tau}.
1048: \label{eq:vfit}
1049: \end{equation}
1050: Table~\ref{tab:times} gives the $\tau$ values derived from this method.
1051: We present results only from $N$-body integrations with $\vk\ge 0.4\ve$
1052: since the smaller kicks did not excite distinct BH-core oscillations,
1053: as discussed above.
1054: To the extent that the motion approximates a
1055: damped SHO, the energy damping time is identical
1056: to the time constant for decay of the kinetic energy alone,
1057: and henceforth we will refer to $\tau$ as the ``energy damping
1058: time constant.''
1059: However in practice, we will use equation~(\ref{eq:vfit})
1060: only to predict changes in $\langle V^2\rangle$.
1061:
1062: The energy damping times in Table~\ref{tab:times} can immediately be
1063: scaled to physical units using equation~(\ref{eq:units}). Such a
1064: scaling presumes that the core properties of our $N$-body models --
1065: which presumably determine $\tau$ -- are related to global properties
1066: in the same way as in real galaxies. A better scheme would relate
1067: $\tau$ directly to the parameters $(\rho_c,\sigma_c,\mbh)$ that
1068: describe the conditions in the core. Since we do not understand the
1069: mechanism(s) responsible for the orbital damping in Phase II, we
1070: experimented with several ways of plotting $\tau$ versus core
1071: parameters.
1072:
1073: Figure~\ref{fig:tauomega} shows that a reasonably tight correlation
1074: exists when $\omega_c\tau$ is plotted against
1075: $\sigma_c^3/(G^{3/2}\rho_c^{1/2}\mbh)$. This is the expected
1076: dependence if dynamical friction is responsible for the damping
1077: (cf. equation~\ref{eq:damp}). However, the effective value of
1078: $\ln\Lambda$ needed to produce the measured damping times is very
1079: small, $0.1 \lesssim\ln\Lambda\lesssim 0.3$
1080: (Figure~\ref{fig:tauomega}). This is yet another way of stating that
1081: orbital decay in Phase II is much slower than predicted by
1082: Chandrasekhar's formula -- roughly a factor $10-20$, if we adopt
1083: $\ln\Lambda\approx 2.5$ for the expected value of the Coulomb
1084: parameter (Fig.~\ref{fig:comp}).
1085:
1086:
1087: \begin{table}
1088: \begin{center}
1089: \caption{Times associated with the evolution in Phases I and II}
1090: \label{tab:times}
1091: \begin{tabular}{cccccccc}
1092: \hline
1093: $\vk/\ve$ & $T_I$ & $\tau$ & $T_{II}$ & &$T_{II}$, &$N_{\rm gal}=$ ... & \\
1094: & & & &$3\times 10^9$&$3\times10^{10}$&$3\times10^{11}$ & $3\times10^{12}$\\
1095: \hline
1096: \multicolumn{8}{c}{A1}\\
1097: \hline
1098: 0.1 & 0.3 & -- & -- & -- & -- & -- & -- \\
1099: 0.2 & 0.3 & -- & -- & -- & -- & -- & -- \\
1100: 0.3 & 0.3 & -- & -- & -- & -- & -- & -- \\
1101: 0.4 & 0.4 & 1.6 & 2.9 & 16.8 & 20.5 & 24.2 & 27.9 \\
1102: 0.5 & 0.7 & 1.3 & 3.2 & 14.5 & 17.5 & 20.5 & 23.4 \\
1103: 0.6 & 1.5 & 1.9 & 4.6 & 21.1 & 25.5 & 30.0 & 34.2 \\
1104: 0.7 & 3.0 & 3.4 & 5.8 & 35.3 & 43.2 & 51.0 & 58.8 \\
1105: 0.8 & 7.3 & 3.8 & 8.5 & 41.6 & 50.3 & 59.0 & 67.8 \\
1106: 0.9 & 20.2 & 2.5 & 6.5 & 28.3 & 34.0 & 39.8 & 45.5 \\
1107: \hline
1108: \multicolumn{8}{c}{A2}\\
1109: \hline
1110: 0.1 & 0.3 & -- & -- & -- & -- & -- & -- \\
1111: 0.2 & 0.3 & -- & -- & -- & -- & -- & -- \\
1112: 0.3 & 0.3 & -- & -- & -- & -- & -- & -- \\
1113: 0.4 & 0.4 & 1.0 & 1.5 & 10.2 & 12.5 & 14.8 & 17.1 \\
1114: 0.5 & 0.7 & 0.95 & 1.9 & 10.2 & 12.4 & 14.5 & 16.7 \\
1115: 0.6 & 1.3 & 1.3 & 3.4 & 14.7 & 17.7 & 20.6 & 23.6 \\
1116: 0.7 & 2.7 & 2.1 & 4.4 & 22.7 & 27.5 & 32.3 & 37.2 \\
1117: 0.8 & 6.5 & 2.4 & 4.6 & 25.4 & 31.0 & 36.5 & 42.0 \\
1118: 0.9 & 20.0 & 2.8 & 6.7 & 31.1 & 37.5 & 43.9 & 50.4 \\
1119: \hline
1120: \multicolumn{8}{c}{B}\\
1121: \hline
1122: 0.1 & 0.5 & -- & -- & -- & -- & -- & -- \\
1123: 0.2 & 0.5 & -- & -- & -- & -- & -- & -- \\
1124: 0.3 & 0.5 & -- & -- & -- & -- & -- & -- \\
1125: 0.4 & 0.55& 2.2 & 3.9 & 23.0 & 28.1 & 33.2 & 38.2 \\
1126: 0.5 & 0.7 & 2.8 & 6.7 & 31.1 & 37.5 & 43.9 & 50.4 \\
1127: 0.6 & 1.3 & 2.6 & 10.8 & 33.4 & 39.4 & 45.3 & 51.4 \\
1128: 0.7 & 2.3 & 2.9 & 10.1 & 35.3 & 42.0 & 48.7 & 55.3 \\
1129: 0.8 & 4.5 & 5.2 & 14.7 & 60.0 & 71.9 & 83.9 & 95.8 \\
1130: 0.9 & 11.5 & 4.3 & 14.9 & 52.3 & 62.2 & 72.1 & 82.0 \\
1131: \hline
1132: \end{tabular}
1133: \end{center}
1134: \end{table}
1135:
1136:
1137: In terms of this scaling, Figure~\ref{fig:tauomega}
1138: allows us to express the damping times in Phase II as
1139: \begin{mathletters}
1140: \begin{eqnarray}
1141: \tau &\approx & 15 \frac{\sigma_c^3}{G^2 \rho_c \mbh}\\
1142: &\approx& 3\times 10^7 {\rm yr}
1143: \left(\frac{\sigma_c}{200\kms}\right)^{-3.86}
1144: \left(\frac{r_c} {30\,{\rm pc}}\right)^2
1145: \label{eq:tauomega}
1146: \end{eqnarray}
1147: \end{mathletters}
1148: where the second line uses the $\mbh-\sigma$ relation \citep{FF:05}.
1149: Based on Figure~\ref{fig:tests} and on the other
1150: arguments given above, we expect the scaling
1151: in equation~(\ref{eq:tauomega}) to be independent of $N$,
1152: i.e. of stellar mass.
1153:
1154: Table~\ref{tab:times} also gives estimates of $T_{II}$, the elapsed
1155: time in Phase II. We defined $T_{II}$ as the time, measured from the
1156: end of Phase I, required for the BH's velocity to fall to its rms
1157: value in Phase III, assuming the time dependence of
1158: equation~(\ref{eq:vfit}). Table~\ref{tab:times} shows that $T_{II}$
1159: is typically longer than $T_I$.
1160:
1161: In a galaxy with $\gg 10^6$ stars, $V_{\rm Brown}^2$ would be much
1162: lower, and $T_{II}$ correspondingly longer, than in our models.
1163: Assuming that the exponential dependence of energy on time persists to
1164: arbitrarily low values of $E$, the additional time spent in Phase II
1165: would be
1166: \begin{equation}
1167: \tau\ln\left(N_{\rm gal}/N\right)
1168: \label{eq:addtime}
1169: \end{equation}
1170: where $N_{\rm gal}$ is the number of stars in the galaxy.
1171: We used the set of $N$-body simulations in Figure~\ref{fig:tests}
1172: to test this dependence.
1173: According to equation~(\ref{eq:addtime}), doubling the number
1174: of particles should extend the elapsed time in Phase II
1175: by an additive amount of $\tau\ln 2 = 0.693\tau\approx 2.2$,
1176: given that the mean $\tau$ value for the four integrations
1177: is $3.2$.
1178: Figure~\ref{fig:tests} confirms this prediction for
1179: $0.25\times 10^6\le N \le 2\times 10^6$.
1180:
1181: Accordingly, Table~\ref{tab:times} also gives values of $T_{II}$
1182: calculated from this formula, for $N_{\rm gal}=(3\times 10^9, 3\times
1183: 10^{10}, 3\times 10^{11}, 3\times 10^{12})$. Conversion from the
1184: $N$-body units of Table~\ref{tab:times} to years is discussed in \S\,\ref{sec:obs}.
1185:
1186: The exponential nature of the damping implies that the distribution
1187: of displacements during Phase II is
1188: approximately uniform in $\ln\Delta r$.
1189:
1190: \section{Effects on the stellar distribution}
1191: \label{sec:profiles}
1192:
1193: The displacement of the BH due to gravitational radiation recoil
1194: affects the stellar distribution and therefore the density profile of
1195: the host galaxy. We expect the stellar structure inside the core to be
1196: particularly affected by the motion of the BH, with important
1197: implications for the shape of the brightness profile in the inner
1198: region. In order to evaluate the changes induced by the escaping BH,
1199: we constructed spatial and projected density profiles for all models
1200: at the end of the simulations, when the BH is well into the Brownian
1201: regime.
1202:
1203: \begin{figure*}
1204: \begin{center}
1205: \includegraphics[width=17cm]{f11.eps}
1206: \end{center}
1207: \caption{Space (left) and projected (right) density profiles
1208: for models A1, A2 and B and different values of the kick
1209: velocity: $\vk = 0.1\,\ve$ (green/dotted), $\vk = 0.3\,\ve$
1210: (red/dashed), $\vk = 0.5\,\ve$ (cyan/dot-dashed), $\vk = 0.9\,\ve$
1211: (blue/long dashed). The black solid lines represent the initial
1212: profile, which is the same for each value of $\vk$.}
1213: \label{fig:density}
1214: \end{figure*}
1215:
1216: Figure~\ref{fig:density} shows the space (left plots) and projected
1217: (right plots) density profiles in models A1, A2, B for $\vk =
1218: (0.1,0.3,0.5,0.9)\,\ve$. We constructed these density profiles using
1219: the kernel-based algorithm of \cite{Merritt:06b}. Particle positions
1220: were first shifted to coordinates that placed the BH at the origin.
1221: The algorithm uses an angle-averaged Gaussian
1222: kernel and modifies the kernel width based on a pilot
1223: (nearest-neighbor) estimate of the density in order to maintain a
1224: roughly constant ratio of bias to variance in the final density
1225: profile. The projected density $\Sigma(R)$ was computed via numerical
1226: projection of the space density. In order to reduce the noise still
1227: further, we combined multiple snapshots at late times and performed
1228: the fit on the combined data sets.
1229:
1230: Figure~\ref{fig:density} shows that a large core develops in the
1231: simulations due to the escape of the BH and its several passages
1232: through the central region. As the BH oscillates under the effect of
1233: the kick, it transfers energy to the surrounding stars, thus pushing
1234: them to larger distances. The stellar density in the core drops and
1235: the slope of the inner distribution decreases, leaving an inner
1236: profile that is flatter than the initial one. The amount of flattening
1237: in the profile or, equivalently, the mass deficit with respect to the
1238: initial profile (shown in the figure with the black solid lines),
1239: increases monotonically with the kick velocity.
1240:
1241: It is interesting to assess whether the final $N$-body profiles are
1242: consistent with the core-S{\'e}rsic law, which is commonly fit
1243: to galaxies with evacuated cores \citep{Graham:03}.
1244: The core-S{\'e}rsic law is:
1245: \begin{mathletters}
1246: \begin{eqnarray}
1247: \Sigma(R) & = &
1248: \Sigma^{'}\left[1+\left(\frac{\rb}{R}\right)^{\alpha}\right]^{\gamma/\alpha}e^{-b
1249: \left[\left(R^{\alpha}+\rb^{\alpha}\right)/\re^{\alpha}\right]^{1/n\alpha}},\\
1250: \Sigma^{'} & = & \Sigma_b \, 2^{-\gamma/\alpha} \, e^{b
1251: \left(2^{1/\alpha}\,\rb/\re \right)^{1/n}}\, ,
1252: \end{eqnarray}
1253: \end{mathletters}
1254: where
1255: $\Sigma_b$ is the density at the break radius $r_b$ and the other parameters are as in
1256: equations~(\ref{eq:PS1}) and~(\ref{eq:PS2}).
1257: To carry out the fits in a manner as similar as possible to
1258: the procedure followed by observers, we counted the projected
1259: particle positions in bins equally spaced in $\log R$.
1260: The parameters $(\re,\rb,\alpha,n,\Sigma^{'})$ were then
1261: varied until the summed residuals in $\mu=-2.5\log\Sigma$
1262: were minimized.
1263:
1264: \begin{table}
1265: \begin{center}
1266: \caption{Fit parameters for models A1, A2 and B.}
1267: \label{tab:fits}
1268: \begin{tabular}{c|rrrrrr}
1269: $\vk$ & $\rb$ & $n$ & $\alpha$ & $\gamma$ & $\re$ & $\Sigma_b$\\
1270: \hline
1271: \multicolumn{7}{c}{A1}\\
1272: \hline
1273: 0.1 & 0.013 & 4.04 & 3.1 & 0.21 & 0.93 & 6.3 \\
1274: 0.2 & 0.016 & 4.05 & 4.2 & 0.24 & 0.93 & 5.6 \\
1275: 0.3 & 0.017 & 4.05 & 4.1 & 0.19 & 0.93 & 5.4 \\
1276: 0.4 & 0.018 & 4.06 & 3.7 & 0.14 & 0.93 & 5.1 \\
1277: 0.5 & 0.019 & 4.06 & 3.6 & 0.11 & 0.93 & 4.9 \\
1278: 0.6 & 0.020 & 4.06 & 3.5 & 0.08 & 0.93 & 4.6 \\
1279: 0.7 & 0.021 & 4.07 & 3.1 & 0.04 & 0.92 & 4.3 \\
1280: 0.8 & 0.022 & 4.07 & 2.8 & 0.02 & 0.92 & 4.1 \\
1281: 0.9 & 0.024 & 4.05 & 2.9 & 0.05 & 0.93 & 3.9 \\
1282: \hline
1283: \multicolumn{7}{c}{A2}\\
1284: \hline
1285: 0.1 & 0.014 & 4.04 & 7.4 & 0.34 & 0.93 & 6.5 \\
1286: 0.2 & 0.015 & 4.04 & 6.7 & 0.31 & 0.93 & 6.2 \\
1287: 0.3 & 0.016 & 4.04 & 6.5 & 0.25 & 0.92 & 5.9 \\
1288: 0.4 & 0.015 & 4.05 & 3.5 & 0.14 & 0.92 & 5.9 \\
1289: 0.5 & 0.017 & 4.05 & 4.0 & 0.12 & 0.92 & 5.4 \\
1290: 0.6 & 0.019 & 4.05 & 4.2 & 0.16 & 0.92 & 4.9 \\
1291: 0.7 & 0.018 & 4.07 & 2.7 & 0.08 & 0.92 & 4.9 \\
1292: 0.8 & 0.022 & 4.06 & 3.6 & 0.07 & 0.92 & 4.4 \\
1293: 0.9 & 0.026 & 4.07 & 4.1 & 0.11 & 0.92 & 3.9 \\
1294: \hline
1295: \multicolumn{7}{c}{B}\\
1296: \hline
1297: 0.1 & 0.020 & 4.05 & 1.9 & 0.16 & 0.92 & 4.4 \\
1298: 0.2 & 0.026 & 4.04 & 2.9 & 0.20 & 0.93 & 3.8 \\
1299: 0.3 & 0.030 & 4.04 & 3.7 & 0.20 & 0.93 & 3.4 \\
1300: 0.4 & 0.034 & 4.05 & 4.3 & 0.16 & 0.92 & 3.1 \\
1301: 0.5 & 0.034 & 4.06 & 3.1 & 0.12 & 0.92 & 3.0 \\
1302: 0.6 & 0.035 & 4.06 & 3.0 & 0.09 & 0.92 & 2.8 \\
1303: 0.7 & 0.039 & 4.08 & 3.0 & 0.07 & 0.91 & 2.6 \\
1304: 0.8 & 0.042 & 4.08 & 2.9 & 0.05 & 0.91 & 2.4 \\
1305: 0.9 & 0.044 & 4.09 & 2.6 & 0.02 & 0.91 & 2.3 \\
1306: \hline
1307: \end{tabular}
1308: \end{center}
1309: \end{table}
1310:
1311: The best-fit parameters for models A1, A2 and B are listed in
1312: Table~\ref{tab:fits}.
1313: \begin{figure}
1314: \begin{center}
1315: \includegraphics[width=8.5cm]{f12.eps}
1316: \end{center}
1317: \caption{Projected density profiles for model A1 computed from the
1318: $N$-body data (points), compared with best-fitting core-S{\'e}rsic
1319: models (lines), for three different
1320: values of the kick velocity ($\vk=0.2,0.4,0.8\,\ve$).
1321: The insert shows a zoom into the central region.}
1322: \label{fig:fits}
1323: \end{figure}
1324: Three of the best fits for model A1 are shown
1325: in Figure~\ref{fig:fits} (lines) together with the projected density
1326: profiles computed from the $N$-body data (points).
1327:
1328: It appears that the host galaxies to recoiling BHs are well
1329: represented by core-S{\'e}rsic profiles. In particular, the fits
1330: show, once again, that the core tends to expand as the BH oscillates
1331: in and out of it, and that the final core size scales as $r_b \sim
1332: \mbh\,\vk^{\beta}$, with $0.3 \simless \beta \simless 0.6$. In
1333: addition, the transition from the inner power law to the outer
1334: S{\'e}rsic profile is rather sharp, with best-fit values of $\alpha$
1335: in the range $2\lesssim \alpha \lesssim 7$. The initial $n=4$ de
1336: Vaucouleurs outer slope is not substantially modified by the BH.
1337:
1338: A flattening of the inner profile is also observed in the simulations of
1339: \citet{Boylan:04}, who follow the evolution of a spherical stellar
1340: bulge with a recoiling central black hole using an $N$-body tree code.
1341: They find that the density profile of the system evolves as a consequence
1342: of the gravitational radiation recoil and flattens substantially.
1343: A core of size equal to the BH sphere of influence forms
1344: on a relatively short time-scale, and remains even after several
1345: dynamical times. A flattening of the profile is observed for
1346: recoil velocities smaller and larger than the central escape speed,
1347: though an additional flattening is present if the black hole returns
1348: to the core after the ejection.
1349:
1350: A measurable signature of a recoiling BH is the mass deficit, the net
1351: mass removed from the central regions \citep{Milos:02}. Mass deficits
1352: produced by recoil will add to the depletion caused by the
1353: pre-existing BH binary, which ejects stars from the core during close
1354: encounters. The deficit produced by the binary is proportional to the
1355: mass of the binary, with only a weak dependence on the mass ratio and
1356: the initial density distribution \citep{Merritt:06}. Therefore, a
1357: binary BH can only produce a deficit $\mdef \approx \mbh$. This could
1358: explain the peak in the distribution of observed mass deficits at
1359: $\mdef/\mbh \approx 1$ \citep{Graham:04,ACS:VI}. The tail of the
1360: distribution, however, extends to values of $\mdef/\mbh \sim 5$.
1361: While such large values might be explained as successive mergers
1362: \citep{Merritt:06}, a recoiling BH represents an interesting
1363: alternative.
1364:
1365: We evaluated the mass deficits in the final $N$-body models by
1366: computing the difference in stellar mass, enclosed within a sphere of
1367: radius $r_s$, between the initial and final space density profiles.
1368: Given the fact that the deficits depend rather sensitively on the
1369: value of $r_s$, we computed $\mdef$ as a function of $r_s$ for a
1370: number of models and kicks. In all cases, $\mdef$ first increases
1371: rapidly with $r_s$ and then flattens out to an approximately constant
1372: value. Based on such tests, we concluded that the most appropriate
1373: values to use for the computation of the mass deficits were as
1374: follows: $r_s = 0.05$ for model A1, $0.04$ for model A2 and $0.1$ for
1375: model B. The results for all three models are shown in
1376: Figure~\ref{fig:mdef}.
1377: \begin{figure}
1378: \begin{center}
1379: \includegraphics[width=8.5cm]{f13.eps}
1380: \end{center}
1381: \caption{Mass deficits, as defined in the text, for the different
1382: runs: A1 (black), A2 (blue), B (red).
1383: Dashed lines show power-law fits.}
1384: \label{fig:mdef}
1385: \end{figure}
1386: Also shown are least-squares fits to $Y=aX^b$, where
1387: $Y\equiv \mdef/\mbh$ and $X\equiv \vk/\ve$.
1388: The best-fit parameters are:
1389: \begin{eqnarray}
1390: {\rm Model\ A1}: a &=&4.83,\ b=1.59 \nonumber \\
1391: {\rm Model\ A2}: a &=&5.08,\ b=1.75 \nonumber \\
1392: {\rm Model\ B\ }: a &=&4.31,\ b=1.90
1393: \end{eqnarray}
1394:
1395:
1396: The largest kicks result in mass deficits as large as $4-5\mbh$, which
1397: is consistent with the largest observed deficits \citep{Merritt:06}.
1398: Our definition of mass deficits as the difference in integrated mass
1399: between initial and final profiles implies that our estimates do not
1400: take into account any depletion prior to the kick. One should
1401: therefore add the contribution from the binary evolution phase ($\mdef
1402: \approx 1 \mbh$) to our measured deficits before comparing with the
1403: observed values.
1404:
1405: The sensitivity of $\mdef$ to $r_s$, which presumably is a feature of
1406: real luminosity profiles as well, suggests that a more objective way
1407: be found to measure mass deficits.
1408:
1409:
1410: We compare the projected density profiles obtained from the $N$-body
1411: simulations to the brightness profiles of a sample of early-type
1412: galaxies in the Virgo cluster observed with the {\it Advanced Camera
1413: for Surveys} (ACS) on the Hubble Space Telescope \citep{ACS:VI}. In
1414: this study, the authors find that, while simple S{\'e}rsic models
1415: generally provide a good representation of the global galaxy profiles,
1416: the brightest galaxies require a power-law component within a
1417: characteristic break radius and are therefore best modeled with
1418: core-S{\'e}rsic profiles.
1419:
1420: We select two representative galaxies in the sample and compare their
1421: surface density profiles with each of the 27 final profiles obtained
1422: from the simulations (9 values of the kick velocities $0.1\dots0.9$
1423: for each of the 3 models A1, A2, B).
1424:
1425: The brightest Virgo galaxy, VCC1226 (M49, NGC 4472), has the largest
1426: value of mass deficit ($\mdef/\mbh \sim 4$) and has a S{\'e}rsic index
1427: that is not too far from our $N$-body models, $n \sim
1428: 5.9$\footnote{Most of the bright galaxies in the ACS sample have $n
1429: \simgreat 7$.}. On the other hand, VCC 731 (NGC 4365) has a relatively
1430: small core and a typical mass deficit of $\sim 1\mbh$
1431: \citep{Merritt:06}. For each galaxy, we scale the $N$-body profiles
1432: to have the same $r_b$ and $\Sigma (r_b)$ as the galaxy itself.
1433:
1434: \begin{figure}
1435: \begin{center}
1436: \includegraphics[width=8.5cm]{f14.eps}
1437: \end{center}
1438: \caption{Surface brightness profiles for the Virgo galaxies VCC 1226
1439: (top) and VCC 731 (bottom) from the ACS sample compared to the
1440: $N$-body profiles obtained from the best fitting of the three
1441: models. The different lines correspond to the 9 different kicks
1442: $\vk/\ve = 0.1\dots0.9$.}
1443: \label{fig:sbp}
1444: \end{figure}
1445: Figure~\ref{fig:sbp} shows that the brightness profiles of both
1446: galaxies can be reasonably well fit by (at least) one of the $N$-body
1447: models. In particular, the profile of VCC 1226 is well fit by models
1448: with $\vk \ge 0.4 \ve \approx 550\kms$ while VCC 731 is well fit by
1449: models with $\vk \simgreat 0.1 \ve \approx 110\kms$.
1450: This indicates that observed brightness profiles, and
1451: even the largest cores, can be well reproduced by the gravitational
1452: recoil kicks.
1453:
1454:
1455: \section{Evolution times in real galaxies}
1456: \label{sec:times}
1457:
1458: Given a galaxy's effective radius $\re$ and total mass $M_{\rm gal}$,
1459: equations~(\ref{eq:units}) relate our $N$-body units to physical
1460: units. We adopt the scaling relations derived from the ACS Virgo
1461: cluster survey of \cite{ACS:I} between $\re$ and absolute blue
1462: magnitude $M_B$ for early-type galaxies. \cite{ACS:VI} found, for
1463: Virgo E galaxies fainter than $M_B\approx -20.5$, a mean relation
1464: \begin{equation}
1465: \log_{10}\re = 0.144 - 0.05 \left(M_B+20\right)
1466: \end{equation}
1467: where $\re$ is in kpc. (Brighter galaxies obey a different relation and are considered
1468: separately below.) We relate $M_B$ to galaxy mass using Gerhard et
1469: al.'s (2001) expression for the mass to light ratio in the blue band:
1470: \begin{equation}
1471: \log_{10}\left[\left(\frac{M}{L}\right)/ \left(\frac{M}{L}\right)_\odot\right]_B \approx 1.17 + 0.67\log_{10}\left(\frac{L_B}{10^{11}\,L_{\odot,B}}\right).
1472: \label{eq:moverl}
1473: \end{equation}
1474: Equation~(\ref{eq:moverl}) was derived from dynamical
1475: modeling of galaxies with $M_B\gtrsim -22.5$
1476: and represents an average for the matter within the effective radius,
1477: including dark matter if present.
1478: Combining these relations gives
1479: \begin{equation}
1480: \re \approx 1.2\ {\rm kpc} \left(\frac{\mgal}{10^{10}\msun}\right)^{0.075}.
1481: \label{eq:revsmb}
1482: \end{equation} The dependence of $\re$ on $\mgal$ is weak, a consequence of the
1483: low slope of the $\re-M_B$ relation. However we note that the scatter
1484: in this relation is large (e.g. \citealp{ACS:VI}, Fig.~136).
1485:
1486: Some fiducial values, and their implied $N$-body scalings
1487: (from equation~\ref{eq:units}), are:
1488: \begin{eqnarray}
1489: \mgal=3\times 10^9\msun\ \ && \ \ \re=1.1\ {\rm kpc} \nonumber \\
1490: &&\left[T\right] = 1.0\times 10^7 {\rm yr} \ \ \ \
1491: \left[V\right] = 110\kms,
1492: \nonumber \\
1493: \mgal=3\times 10^{10}\msun\ \ && \ \ \re=1.3\ {\rm kpc} \nonumber \\
1494: &&\left[T\right] = 4.1\times 10^6 {\rm yr} \ \ \ \
1495: \left[V\right] = 315\kms,
1496: \nonumber \\
1497: \mgal=3\times 10^{11}\msun\ \ && \ \ \re=1.5\ {\rm kpc} \nonumber \\
1498: &&\left[T\right] = 1.7\times 10^6 {\rm yr} \ \ \ \
1499: \left[V\right] = 910\kms.
1500: \nonumber
1501: \label{eq:scalings}
1502: \end{eqnarray}
1503: The trend of decreasing $\left[T\right]$ with increasing $\mgal$
1504: reflects the well-known higher density of more massive galaxies
1505: \citep{Graham:03}. The central escape velocities in our models are
1506: $2.0\lesssim\ve\lesssim 2.2$ in $N$-body units, corresponding to
1507: $\ve\approx 2.1\times\left[V\right] \approx 2000\kms$ when scaled to a
1508: $3\times 10^{11}\msun$ galaxy. This agrees well with escape
1509: velocities of bright E-galaxies derived from more detailed modeling
1510: (e.g. Fig.~2, \citealp{MMFHH:04}).
1511:
1512: All of the times listed in Table~\ref{tab:times} can be scaled to
1513: physical units using these relations. Figure~\ref{fig:times} shows
1514: the result for the three fiducial values of $\mgal$. We have used
1515: equation~(\ref{eq:addtime}) to correct the measured $T_{II}$ values to
1516: different values of $N_{\rm gal}\equiv \mgal/m_\star$ assuming
1517: $m_\star=\msun$; we also show, as conservative lower limits, the
1518: $T_{II}$ times obtained directly from the simulations.
1519:
1520:
1521: \begin{figure}
1522: \begin{center}
1523: \includegraphics[width=7.cm]{f15a.eps}
1524: \includegraphics[width=7.cm]{f15b.eps}
1525: \includegraphics[width=7.cm]{f15c.eps}
1526: \end{center}
1527: \label{fig:times1}
1528: \caption{Return times of kicked BHs.
1529: These are plots of the $N$-body values given in
1530: Table~\ref{tab:times}, scaled to physical units
1531: using equations~(\ref{eq:units}) and~(\ref{eq:revsmb}).
1532: {\it Lower (filled) symbols:} $T_I$;
1533: {\it Middle (open) symbols:} $T_I+T_{II}$, with $T_{II}$ taken
1534: directly from the simulations;
1535: {\it Upper (filled) symbols:} $T_I+T_{II}$, with $T_{II}$
1536: scaled to $N_{\rm gal}$ using equation~(\ref{eq:addtime}).
1537: (a) $\mgal=3\times 10^9\msun$;
1538: (b) $\mgal=3\times 10^{10}\msun$;
1539: (c) $\mgal=3\times 10^{11}\msun$.
1540: }
1541: \label{fig:times}
1542: \end{figure}
1543:
1544:
1545: Figure~\ref{fig:times} seem to suggest that return times depend
1546: discontinuously on $\vk$, since Phase II does not appear to exist in
1547: our simulations when $\vk\lesssim 0.3\ve$. As discussed above, this might
1548: not be true in simulations with larger $N$, or in real galaxies. In
1549: any case, for $\vk\gtrsim 0.4\ve$, return times are dominated by the
1550: time spent in Phase II (``BH-core oscillations'').
1551:
1552: The brightest galaxies, $M_B\lesssim -21$, appear to obey a different
1553: scaling relation between $\re$ and $M_B$ than the
1554: relation~(\ref{eq:revsmb}) given above \citep{ACS:VI}. Furthermore,
1555: these bright galaxies are typically fit by S{\'e}rsic indices in the
1556: range $5\lesssim n \lesssim 10$, larger than the value $n=4$ adopted
1557: here for the $N$-body models. On the other hand, the brightest E
1558: galaxies often have resolved cores with well-determined sizes and
1559: densities (cf. \S\,\ref{sec:profiles}). Furthermore,
1560: equation~(\ref{eq:tauomega}) gives the damping time $\tau$ in Phase II
1561: in terms of core properties alone.
1562:
1563: We define the Phase II return times for these galaxies
1564: as the time for the BH's energy to decrease from
1565: \begin{equation}
1566: \frac{1}{2} \omega_c^2 r_c^2 \approx \frac{2}{3} \pi G \rho_c r_c^2,
1567: \end{equation}
1568: the BH's energy when it first re-enters the core, to
1569: \begin{equation}
1570: \frac{1}{2} V_{\rm brown}^2 \approx \frac{3}{2}\frac{m_\star}{\mbh}\sigma_c^2,
1571: \end{equation}
1572: the Brownian energy,
1573: assuming an energy damping time constant of $\tau$;
1574: for the latter we take equation~(\ref{eq:tauomega}).
1575: This time is
1576: \begin{mathletters}
1577: \begin{eqnarray}
1578: T_{II} &=& \cal{N} \tau, \nonumber \\
1579: \tau &\approx & 15 \frac{\sigma_c^3}{G^2 \rho_c\,\mbh}
1580: \label{eq:vt1} \\
1581: &\approx &1.2\times 10^7{\rm yr}
1582: \left(\frac{\sigma_c}{250\kms}\right)^3
1583: \left(\frac{\rho_c}{10^3\msun\,{\rm pc}^{-3}}\right)^{-1}
1584: \left(\frac{\mbh}{10^9\msun}\right)^{-1}, \nonumber \\
1585: \cal{N} &=& \ln\left(\frac{1}{F^2}\frac{\mbh}{m_\star}\right) \nonumber \\
1586: &\approx& \ln\left(\frac{1}{F^2} \frac{\mbh}{\mgal} \frac{\mgal}{\msun}\right),
1587: \label{eq:vt2}
1588: \end{eqnarray}
1589: \end{mathletters}
1590: with $F\approx 2$ the form factor defined above,
1591: and we have again assumed $m_\star=\msun$.
1592:
1593: \begin{figure}
1594: \begin{center}
1595: \includegraphics[width=8.5cm]{f16.eps}
1596: \end{center}
1597: \caption{Estimates of the return time in Phase II for supermassive
1598: black holes in the six brightest Virgo galaxies, excluding
1599: M87 \citep{ACS:I}.
1600: This plot assumes that the SMBHs have received a kick
1601: large enough to remove them from the core initially.
1602: Lower (open) symbols show the energy decay time constant
1603: in the core, $\tau$ (equation~\ref{eq:vt1}),
1604: while upper (filled) symbols show $\cal{N}\tau$,
1605: where $\cal{N}$ is the estimated number of time constants
1606: required for the BH's velocity
1607: to decay to the Brownian value (equation~\ref{eq:vt2}).
1608: }
1609: \label{fig:virgotimes}
1610: \end{figure}
1611:
1612: Figure~\ref{fig:virgotimes} shows estimates of $\tau$ and
1613: $\cal{N}\tau$ for the six brightest galaxies in the ACS Virgo sample
1614: excluding M87, which has an active nucleus \citep{ACS:I}. Of course,
1615: this figure is only meaningful under the assumption that the BHs in
1616: these galaxies have received large enough kicks to remove them
1617: completely from the core, i.e. $\vk\approx 10^3\kms$. But if this did
1618: occur, Figure~\ref{fig:virgotimes} suggests that return times would be
1619: of order $1$\,Gyr. Such a long time is comparable with the mean time
1620: between galaxy mergers in a dense environment like the Virgo cluster.
1621: Hence, a SMBH might never return fully to the center before another
1622: SMBH spirals in.
1623:
1624: \section{Observable Consequences}
1625: \label{sec:obs}
1626:
1627: \subsection{Likelihood of Large Kicks}
1628: \label{sec:kicks}
1629: Kicks large enough to remove SMBHs from cores,
1630: $\vk\gtrsim 0.4\ve$, range from
1631: $\sim 90\kms$ for $\mgal=3\times 10^9\msun$,
1632: to $\sim 750\kms$ for $\mgal=3\times 10^{11}\msun$,
1633: to $\sim 1000\kms$ for $\mgal=3\times 10^{12}\msun$,
1634: based on the fiducial scalings in \S\,\ref{sec:times}.
1635: The most propitious configuration for the kicks appears to be
1636: an equal-mass binary in which the individual
1637: spin vectors are oppositely aligned and oriented parallel to the
1638: orbital plane \citep{Campanelli:07b,Gonzalez:07b}.
1639: Assuming this most favorable orientation,
1640: and setting the spins to their maximal values,
1641: the maximum kick (oriented parallel to the binary angular
1642: momentum vector) is believed to scale with binary mass ratio
1643: $q\equiv M_2/M_1\le 1$ as
1644: \begin{equation}
1645: V_{\rm max} \approx 6\times 10^4 {\rm km\ s}^{-1} {q^2\over (1+q)^4}
1646: \end{equation}
1647: \citep{Lousto:07}.
1648: Mass ratios as small as $q\approx 0.2$ can therefore
1649: result in kicks $\gtrsim 1000$ km s$^{-1}$.
1650: While the assumption of near-maximal spins is probably not an extreme
1651: one (e.g. \citealp{Shapiro:05,Gammie:04}),
1652: orienting the BHs with their spins perpendicular to the orbital
1653: angular momentum may seem odd, particularly in gas-rich galaxies
1654: \citep{BRM:07}.
1655: However there is considerable circumstantial evidence that SMBH
1656: spin axes bear no relation
1657: to the orientations of the gas disks that surround them
1658: \citep{Kinney:00,Gallimore:06,Borguet:07}
1659: and this is presumably even more true with
1660: respect to the directions of infalling BHs in gas-free galaxies.
1661: If SMBH spins do orient parallel with orbital angular momenta,
1662: the maximum kick is more modest and contains contributions
1663: from both the ``mass asymmetry'' ($M_1\ne M_2$) and from the
1664: spins.
1665: The two kick components, both of which are parallel to the orbital plane,
1666: are believed to be approximately independent
1667: and to scale roughly as
1668: \begin{mathletters}
1669: \begin{eqnarray}
1670: V_{\rm mass} &\approx& V_1{q^2(1-q)\over (1+q)^5}, \\
1671: V_{\rm spin} &\approx& V_2{q^2\over (1+q)^5}\left(\alpha_2 - q\alpha_1\right);
1672: \label{eq:V0}
1673: \end{eqnarray}
1674: \end{mathletters}
1675: $V_1\approx V_2\approx 10^4$ km s$^{-1}$ and
1676: $\alpha$ denotes a dimensionless spin, $-1\le\alpha\le 1$
1677: \citep{Lousto:07,Baker:07,LZ:07}.
1678: $V_{\rm mass}$ peaks at $\sim 200$ km s$^{-1}$
1679: for $q\approx 0.4$ while $V_{\rm spin}$ peaks at $\sim 600$ km s$^{-1}$
1680: for $q\approx 1, \alpha_1=-\alpha_2=1$.
1681: In this less-favorable configuration, kicks could remove
1682: SMBHs only from the cores of low-to-moderate luminosity galaxies.
1683: Estimates of the kick velocity distribution (e.g. \citealp{SB:07})
1684: are extremely uncertain since they depend on the unknown distributions
1685: of SMBH mass ratios, spins and spin orientations.
1686: In what follows, we will focus on the consequences of kicks
1687: that are large enough to remove SMBHs from galaxy cores
1688: and to excite the long-lived oscillations that we described
1689: above.
1690:
1691: \subsection{Offset and Double Nuclei}
1692: \label{sec:nuclei}
1693:
1694: \cite{Lauer:05} identified five galaxies in which the point
1695: of maximum surface brightness is displaced from the
1696: center of the isophotes defined by the galaxy on large scales.
1697: All are luminous, ``core'' galaxies.
1698: Contour plots for two of the galaxies, NGC 507 and 1374
1699: (Figs.~17, 18 of \citealt{Lauer:05}), look strikingly similar
1700: to the ``Phase II'' isodensity plots in Figure~\ref{fig:cont}.
1701: Displacements are cited for NGC 507 ($0''.06\approx 19\rm pc$),
1702: NGC 1374 ($0''.02\approx 2.1\rm pc$), and
1703: NGC 7619 ($0''.04\approx 11\rm pc$), all of which are of
1704: order the core radii in these galaxies.
1705: The five galaxies with offset nuclei comprise 12\% of the
1706: Lauer et al. ``core'' galaxy sample;
1707: no offset nuclei were found among the ``power-law''
1708: (non-cored) galaxies.
1709: Several of the offsets are close to the resolution
1710: limit, and some offsets might go unobserved due to
1711: projection, so it is likely that offset nuclei are
1712: quite common in ``core'' galaxies.
1713: If the offsets are produced by oscillations like those
1714: in Figure~\ref{fig:cont}, the SMBHs in these galaxies would be
1715: located on the opposite side of the galaxy photocenter
1716: from the point of peak brightness.
1717: Phase II oscillations can also produce a ``double nucleus''
1718: morphology (e.g. Figure~\ref{fig:cont}, frame 8) with the BH
1719: located at either the higher or secondary peak.
1720: This is a reasonable model for the double nucleus in NGC 4486B
1721: \citep{Lauer:96}, since the two peaks are closely matched
1722: in brightness and are offset by similar amounts ($\sim 6\rm pc$)
1723: from the galaxy photocenter.
1724: Galaxies with central minima in the surface brightness
1725: (e.g. NGC 4406, NGC 6876; \cite{Lauer:02}) might also be
1726: explained in this way.
1727: This model is probably not as appropriate for the more famous
1728: double nucleus in M31, since M31 is not a ``core'' galaxy,
1729: and one of the brightness peaks (the one associated
1730: with the SMBH) lies close to the galaxy photocenter \citep{Lauer:93}.
1731:
1732: \subsection{Displaced AGN}
1733: \label{sec:AGN}
1734:
1735: An ejected SMBH can appear as a spatially or kinematically
1736: displaced AGN \citep{Kapoor:76,Kapoor:83a,Kapoor:83b}.
1737: A recoiling SMBH retains gas that is orbiting around it
1738: within a distance
1739: \begin{equation}
1740: r_{\rm eff} \approx \frac{G \mbh}{\vk^2} \approx
1741: 0.5\,{\rm pc}\ M_8 V_{\rm k, 1000}^{-2}
1742: \end{equation}
1743: with $M_8\equiv\mbh/10^8\msun$ and $V_{\rm k, 1000}\equiv \vk/1000\kms$.
1744: An accretion disk if present would mostly be retained,
1745: and for kicks $\lesssim 10^3\kms$, $r_{\rm eff}$ is large enough
1746: to encompass most of the broad emission-line region gas as well.
1747: Narrow emission lines originate in gas moving in the gravitational
1748: potential of the host galaxy and would not follow a recoiling SMBH
1749: \citep{Merritt:HE0450}.
1750: \cite{Bonning:07} used this argument to search for kinematic
1751: offsets between spectral features associated with the
1752: broad- and narrow emission line regions.
1753: No convincing cases were found.
1754: This may be a consequence of the rapid decrease in SMBH energy
1755: during Phase I (Fig.~\ref{fig:eoft}).
1756: In the longer-lived oscillations that characterize Phase II,
1757: the rms velocity of the SMBH drops from
1758: \begin{equation}
1759: \sim 90\kms
1760: \left(\frac{\rho_c}{10^3\msun\ {\rm pc}^{-3}}\right)^{1/2}
1761: \left(\frac{r_c}{30\,{\rm pc}}\right)
1762: \end{equation}
1763: when it first re-enters the core, to
1764: \begin{equation}
1765: \sim 0.03\kms \left(\frac{\mbh}{10^8\msun}\right)^{-1/2}
1766: \left(\frac{\sigma_c}{200\kms}\right)
1767: \end{equation}
1768: in the Brownian regime.
1769: Such small velocity offsets would be difficult to detect.
1770: An alternative approach would be to search for linear
1771: displacements $\Delta R$ between the AGN emission and the
1772: peak of the stellar surface brightness.
1773: This displacement is $\sim r_c$ at the start of Phase II,
1774: dropping to $\sim \sqrt{m_\star/\mbh}r_c$ in the Brownian
1775: regime; the exponential nature of the damping implies
1776: an approximately uniform distribution of
1777: $\ln\Delta R$ during Phase II.
1778: Relatively large ($\sim 10-100 \rm pc$) offsets between the AGN and either the
1779: stellar density peak or the center of rotation have in fact
1780: been claimed in a number of galaxies based on integral-field
1781: spectroscopy \citep[e.g.][]{Media:93,Media:05}.
1782:
1783: \subsection{Wiggling Jets}
1784: \label{sec:jets}
1785:
1786: During Phase II, the SMBH oscillates sinusoidally within the
1787: core with roughly constant period,
1788: \begin{equation}
1789: \frac{2\pi}{\omega_c} \approx 1.4\times 10^6 \,{\rm yr}
1790: \left(\frac{\rho_c}{10^3\msun\ {\rm pc}^{-3}}\right)^{-1/2},
1791: \end{equation}
1792: and with velocities as given above.
1793: Such motion will induce periodic deviations in the
1794: velocity and direction of a jet emitted by the SMBH
1795: \citep{Roos:92}.
1796: If the jet is oriented perpendicularly to the direction of
1797: motion of the SMBH, the jet direction is fixed, and the
1798: jet material moves on a cylindrical surface with radius
1799: equal to the radius of the SMBH's orbit.
1800: If the jet velocity has some component parallel to the
1801: SMBH's motion, the two velocities add and the cylinder becomes
1802: a cone over which the jet precesses \citep{Roos:93}.
1803: Such models have been used to explain the helical distortions
1804: observed in a number radio sources;
1805: the inferred orbital periods are typically $1-100\,\rm yr$,
1806: and the jet accelerations are usually ascribed to the orbit of the
1807: jet-producing SMBH around a second SMBH in a close
1808: ($\ll 1 \rm pc$) binary pair.
1809: However some sources are fit by models with longer periods.
1810: For instance, the morphology of the {\bf C}-type source
1811: 3C 449 has been reproduced assuming jet forcing
1812: with a period of $\sim 10^7\,\rm yr$ \citep{Hardee:94}.
1813: Such long periods are sometimes explained in terms of
1814: bulk motion of the galaxy hosting the radio source
1815: \citep{Icke:78},
1816: but oscillations of the SMBH within the core might
1817: provide a tenable alternative in some cases.
1818:
1819:
1820: \subsection{Oversized Cores and Hypermassive Black Holes}
1821: \label{sec:hyper}
1822:
1823: Cores generated by kicked SMBHs can be substantially larger
1824: than those produced by ``core scouring'' from a binary SMBH
1825: \citep{MM:01,Merritt:06}, particularly when
1826: $\vk\gtrsim 0.4\ve$.
1827: As shown in \S\,\ref{sec:profiles} (Figures~\ref{fig:density}-\ref{fig:sbp}),
1828: kick-induced cores can be as large as those
1829: observed in some of the brightest ``core'' galaxies,
1830: having mass deficits of $4-5\mbh$ and core radii
1831: several times the SMBH influence radius,
1832: or $\sim 5\%$ of the galaxy's half-light radius.
1833: (Similar conclusions were reached already by
1834: \cite{Boylan:04} and \cite{MMFHH:04}.)
1835: While the majority of observed mass deficits lie in the
1836: range $0.5\lesssim \mdef/\mbh\lesssim 1.5$,
1837: some E galaxies have $\mdef/\mbh\gtrsim 3$,
1838: too large to be easily explained by core scouring.
1839: \cite{Lauer:07} invoked the oversized cores, along with other
1840: circumstantial evidence, to argue that the SMBHs in the brightest
1841: E galaxies are ``hypermassive,'' $\mbh\gtrsim 10^{10}\msun$.
1842: An alternative possibility is that the largest cores have been
1843: enlarged by kicks.
1844: Figure~\ref{fig:mdef}, combined with earlier $N$-body results
1845: \citep{Merritt:06}, suggests that the total mass deficit
1846: generated by a binary SMBH following a single galaxy merger is
1847: \begin{eqnarray}
1848: \mdef &=&M_{\rm def,bin} + M_{\rm def,kick} \nonumber \\
1849: M_{\rm def,bin} &\approx& 0.7 q^{0.2}\mbh, \nonumber \\
1850: M_{\rm def,kick} &\approx& 5\mbh \left(\vk/\ve\right)^{1.75}
1851: \label{eq:mdefofv}
1852: \end{eqnarray}
1853: where $M_{\rm def,bin}$ and $M_{\rm def,kick}$ are the mass deficits
1854: generated by ``core scouring'' and by the kick respectively and
1855: $q\equiv M_2/M_1\le 1$ is the binary mass ratio.
1856: It has been argued \citep{Merritt:06} that the ratio
1857: $M_{\rm def,bin}/\mbh$ increases in multiple mergers,
1858: and the same is likely to be
1859: true for kick-induced core growth.
1860: Thus, the decrease in typical values of of $\vk/\ve$ with
1861: increasing galaxy luminosity might be offset by the greater
1862: number of mergers that contribute to the growth of luminous
1863: galaxies, leading to comparable values of $\mdef/\mbh$.
1864: In any case, the possibility that core growth is dominated
1865: by the kicks should be considered in future studies.
1866:
1867: \acknowledgements
1868: We thank D. Axon, M. Campanelli, S. Komossa, C. Lousto, R. Miller,
1869: A. Robinson and Y. Zlochower for illuminating discussions. This work
1870: was supported by grants AST-0206031, AST-0420920 and AST-0437519 from
1871: the NSF, grant NNG04GJ48G from NASA, and grant HST-AR-09519.01-A from
1872: STScI.
1873:
1874: \bibliographystyle{apj}
1875: \bibliography{biblio}
1876:
1877: \end{document}
1878: