0704.1968/ms.tex
1: %\documentclass[12pt,preprint]{aastex}
2: % version of June 23, 2008 - London
3: % gergely@physx.u-szeged.hu
4: 
5: 
6: \documentclass[12pt,preprint]{aastex}
7: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
8: \usepackage{amssymb}
9: 
10: \def\AJ{{\it Astron. J.} }
11: \def\ARAA{{\it Annual Rev. of Astron. \& Astrophys.} }
12: \def\ApJ{{\it Astrophys. J.} }
13: \def\ApJL{{\it Astrophys. J. Letters} }
14: \def\ApJS{{\it Astrophys. J. Suppl.} }
15: \def\ApP{{\it Astropart. Phys.} }
16: \def\AA{{\it Astron. \& Astroph.} }
17: \def\AAR{{\it Astron. \& Astroph. Rev.} }
18: \def\AAL{{\it Astron. \& Astroph. Letters} }
19: \def\JGR{{\it Journ. of Geophys. Res.}}
20: \def\JPhG{{\it Journ. of Physics} {\bf G} }
21: \def\PhFl{{\it Phys. of Fluids} }
22: \def\PR{{\it Phys. Rev.} }
23: \def\PRD{{\it Phys. Rev.} {\bf D} }
24: \def\PRL{{\it Phys. Rev. Letters} }
25: \def\Nature{{\it Nature} }
26: \def\MNRAS{{\it Month. Not. Roy. Astr. Soc.} }
27: \def\Science{{\it Science} }
28: \def\ZA{{\it Zeitschr. f{\"u}r Astrophys.} }
29: \def\ZFN{{\it Zeitschr. f{\"u}r Naturforsch.} }
30: \def\etal{{\it et al.}}
31: \shorttitle{Supermassive black hole binary mergers} 
32: \shortauthors{L.\'{A}. Gergely and P.L. Biermann} 
33: 
34: \begin{document}
35: 
36: \title{The spin-flip phenomenon \\
37: in supermassive black hole binary mergers}
38: \author{L\'{a}szl\'{o} \'{A}rp\'{a}d Gergely$^{1,2,3\star }$ and Peter L.
39: Biermann$^{4,5,6,7,8\ddag }$}
40: 
41: 
42: \affil{
43: $^{1}$Department of Theoretical Physics, University of Szeged, Tisza Lajos krt 84-86, 
44: Szeged 6720, Hungary\\
45: $^{2}$Department of Experimental Physics, University of Szeged, D\'{o}m t\'{e}r 9, 
46: Szeged 6720, Hungary\\
47: $^{3}$Department of Applied Science, London South Bank University, 103 Borough Road, 
48: London SE1 0AA, UK\\
49: $^{4}$Max Planck Institute for Radioastronomy, Bonn, Germany\\
50: $^{5}$Department of Physics and Astronomy, University of Bonn, Germany\\
51: $^{6}$Department of Physics and Astronomy, University of Alabama,
52: Tuscaloosa, AL, USA \\
53: $^{7}$Department of Physics, University of Alabama at Huntsville, AL, USA 
54: \\
55: $^{8}$FZ Karlsruhe and Physics Department, University of Karlsruhe, Germany\\
56: \small {$^\star$ E-mail: gergely@physx.u-szeged.hu\qquad $^\ddag$ E-mail:
57: plbiermann@mpifr-bonn.mpg.de} }
58: 
59: 
60: \begin{abstract}
61: Massive merging black holes will be the primary sources of powerful
62: gravitational waves at low frequency, and will permit to test general
63: relativity with candidate galaxies close to a binary black hole merger. In
64: this paper we identify the typical mass ratio of the two black holes but
65: then show that the distance where gravitational radiation becomes the
66: dominant dissipative effect (over dynamical friction) does not depend on the
67: mass ratio; however the dynamical evolution in the gravitational wave
68: emission regime does. For the typical range of mass ratios the final stage
69: of the merger is preceded by a rapid precession and a subsequent spin-flip
70: of the main black hole. This already occurs in the inspiral phase, therefore
71: can be described analytically by post-Newtonian techniques. We then identify
72: the radio galaxies with a superdisk as those in which the rapidly precessing
73: jet produces effectively a powerful wind, entraining the environmental gas
74: to produce the appearance of a thick disk. These specific galaxies are thus
75: candidates for a merger of two black holes to happen in the astronomically
76: near future.
77: \end{abstract}
78: 
79: \keywords{compact binaries, gravitational radiation, radio galaxies,jets}
80: 
81: \label{firstpage}
82: 
83: \section{Introduction}
84: 
85: The most energetic phenomenon that involves general relativity in the
86: observable universe is the merger of two supermassive black holes (SMBHs).
87: Therefore the study of these mergers may provide one of the most stringent
88: tests of general relativity even before the discovery and precise
89: measurement of the corresponding gravitational waves (see, e.g., Sch{\"{a}}%
90: fer 2005).
91: 
92: Most galaxies have a central massive black hole (Kor\-mendy \& Richstone
93: 1995, Sanders \& Mirabel 1996, Faber et al. 1997), and after their initial
94: growth (for one possible example how this might happen, see Munyaneza \&
95: Biermann, 2005, 2006), their evolution is governed by mergers. Therefore,
96: the two central black holes also merge (see Zier \& Biermann 2001, 2002;
97: Biermann et al. 2000; Merritt \& Ekers 2002; Merritt 2003; Gopal-Krishna et
98: al. 2003, 2004, 2006; Gopal-Krishna \& Wiita 2000, 2006; Zier 2005, 2006,
99: 2007). Before the two black holes get close, the galaxies begin to round
100: each other, distorting the shape of a radio galaxy fed by one or both of the
101: two black holes; thence the Z-shaped radio galaxies (Gopal-Krishna et al.
102: 2003). When they merge, under specific circumstances to be clarified in this
103: paper, a spin-flip may occur. For a black hole nurturing activity around it,
104: the spin axis defines the axis of a relativistic jet, and therefore a
105: spin-flip results in a new jet direction: thence the X-shaped radio galaxies
106: (Rottmann 2001, Chirvasa 2001, Biermann et al. 2000, Merritt and Ekers
107: 2002). In fact, observations suggest that all activity around a black hole
108: may result in a relativistic jet even for radio-weak quasar activity (Falcke
109: et al. 1996, Chini et al. 1989a, b). A famous color picture showing the past
110: spin-flip of the M87 black hole (Owen et al. 2000) clearly shows a weak
111: radio counter-jet, misaligned with the modern active jet by about 30$^{\circ
112: }$. The feature of the X-shaped radio galaxy jets is so common and yet very
113: short-lived that all radio galaxies may have been through this merger
114: (Rottmann 2001), and thus should have undergone a spin-flip. This can be
115: also deduced from the observation that many compact steep spectrum sources
116: show a misaligned double radio structure, where an inner pair of hot spots
117: is misaligned with an outer pair of hot spots (Marecki et al. 2003). We
118: conclude that theoretical arguments and observations consistently suggest
119: that black holes merge and result in a spin-flip.
120: 
121: From these and some other data we deduce a few basic tenets that the theory
122: needs to explain:
123: 
124: 1. In the X-shaped radio galaxies the angles between two pairs of jets in
125: projection are typically less than 30 degrees. The real angles can be even
126: about 45$^{\circ }$. The jets are believed to signify the spin axis of the
127: more active (therefore presumably the more massive) black hole before the
128: merger and the spin axis of the merged black hole. Therefore, a substantial
129: spin-flip should have occurred.
130: 
131: 2. In the X-shaped radio galaxies one pair of jets has a steep radio
132: spectrum. This implies that it has not recently been resupplied
133: energetically, it is an old pair of jets; and its synchrotron age is
134: typically a few $10^{7}$ years. The other pair of jets has a relatively flat
135: radio spectrum (this is the new jet; Rottmann 2001). Radio continuum
136: spectroscopy thus supports the spin-flip model.
137: 
138: 3. Again, as Rottmann (2001) shows, the statistics of X-shaped radio
139: galaxies are such, that every radio galaxy may have passed through this
140: stage during its evolution. This matches with arguments based on
141: far-infrared observations that central activity in galaxies such as
142: starbursts and feeding of the activity of a central black hole, is often,
143: maybe always, preceded by a merger of galaxies (Sanders \& Mirabel 1996).
144: 
145: 4. There is another critical observation of the spectrum of radio galaxies.
146: For many of them the radio spectrum has a low-frequency cutoff, suggesting a
147: cutoff in the energy distribution of the electrons at approximately the pion
148: mass (the electrons / positrons are decay products from pions, produced in
149: hadronic collisions; Falcke et al. 1995, Biermann et al. 1995, Falcke and
150: Biermann 1995a, 1995b, 1999, Gopal-Krishna et al. 2004). Hadronic collisions
151: with ensuing pion production at the foot ring of the radio jet occur
152: naturally and thermally in the case that the rotation parameter of the black
153: hole is larger than 0.95, and if the foot ring is an advection-dominated
154: accretion flow (ADAF) or radiatively inefficient accretion flow (RIAF; Donea
155: \& Biermann 1996, Mahadevan 1998, Gopal-Krishna et al. 2004). If this is
156: true for all radio galaxies, the spin of the black hole both before and
157: after the spin-flip must be more than $95~\%$ of the maximally allowed
158: value. This is a major constraint on the process of the spin-flip. If we
159: assume that this holds for all radio galaxies, then a fortiori it also holds
160: for those which have just undergone a binary black hole merger, and so their
161: spin ought to be high as well.
162: 
163: 5. When two black holes merge, the emission of strong gravitational waves is
164: certain (Peters \& Mathews 1963, Peters 1964, Thorne 1979). Compact binaries
165: are driven by gravitational radiation through a post-Newtonian (PN) regime
166: (the inspiral), a plunge and a ring-down phase toward the final state. It is
167: commonly believed than the spin-flip phenomenon is likely to be caused by
168: the gravitational radiation escaping the merging system (Rottmann 2001,
169: Biermann et al. 2000, Merritt and Ekers 2002). Recent numerical work on the
170: final stages of the coalescence supports this (see Br{\"{u}}gmann 2007;
171: Campanelli et al. 2007a, b; Gonzalez et al. 2007a, b).
172: 
173: Therefore, it is mandatory to investigate what happens when the two black
174: holes get close to each other, and this we propose to treat in this paper.
175: We present here a model which allows \textit{to have a merger transition
176: going from a high-spin stage to another high-spin stage}, using mostly
177: physical insight from outside of the innermost stable orbit (ISO). In
178: contrast with available numerical simulation, our method, limited to a
179: certain typical range of mass ratios of the two black holes, has the
180: advantage that the evolution of the compact binary can be treated in the
181: framework of an analytical PN expansion with two small parameters.
182: 
183: In Section 2, we review the current state of observations on the masses of
184: supermassive galactic black holes, which roughly scale with the bulge masses
185: of their host galaxies. The observations suggest that the most massive black
186: holes have about $3\times 10^{9}$ solar masses ($M_{\odot }$) and the most
187: reliable determination of the low-mass central black hole (in our galaxy) is
188: about $3\,\times 10^{6}$ $M_{\odot }$ (Ghez et al. 2005, Sch\"{o}del \&
189: Eckart 2005). There is some evidence for central massive black holes of
190: slightly lower mass (Barth et al. 2005), but the error bars are very large.
191: This implies that the maximum mass ratio is about $10^{3}$. We carefully
192: analyze the statistics and argue in Section 2 that mass ratios in the range $%
193: 3:1$ to $30:1$ cover most of the plausible range in mergers of galactic
194: central black holes. Roughly speaking, this means that typically one mass is
195: dominant by a factor of order 10. Therefore, we find that neither the much
196: discussed case of equal masses nor that of the extreme mass ratios (test
197: particles falling into a black hole) describes typical central galactic SMBH
198: mergers.
199: 
200: In Section 3, we study the relative magnitudes of the spin of the dominant
201: black hole and of the orbital angular momentum of the system. Their ratio
202: depends on two factors: the mass ratio and the separation of the binary
203: components (the inverse of which scales with the post-Newtonian parameter).
204: We show that for the typical mass ratio interval the orbital momentum left
205: when the system is reaching ISO is much smaller than the dominant spin. So
206: in the typical mass range case whatever happens during the plunge and
207: ring-down phases of the merger, in which the remaining orbital momentum is
208: dissipated, it cannot change essentially the direction of the spin. By
209: contrast, for equal mass mergers the orbital angular momentum dominates
210: until the end of the inspiral, while for extreme mass ratio mergers the
211: larger spin dominates from the beginning of the gravitational wave driven
212: merger phase.
213: 
214: In Section 4, we discuss the transition from the dynamical friction
215: dominated regime to the gravitational radiation dominated regime, in order
216: to establish the initial data for the PN treatment. The interaction of the
217: black holes with the already merged stellar environment generates a
218: dynamical friction when the separation of the black holes is between a few
219: parsecs (pc) and one hundredth of a pc. Gravitational radiation has a small
220: effect in this regime. Due to the dynamical friction, some of the orbital
221: angular momentum of the binary black hole system is transferred to the
222: stellar environment, such that the stellar population at the poles of the
223: system tends to be ejected and a torus is formed (Zier \& Biermann 2001,
224: 2002, Zier 2006). This connects to the ubiquitous torus around active
225: galactic nuclei (AGNs), detected first in X-ray absorption (Lawrence \&
226: Elvis 1982, Mushotzky 1982), and later confirmed by optical polarization of
227: emission lines (Antonucci \& Miller 1985). Dynamical friction is enhanced as
228: in a merger the phase-space distribution is strongly disturbed by large
229: fluctuations of the mass distribution (Lynden-Bell 1967, Toomre \& Toomre
230: 1972, Barnes \& Hernquist 1992, Barnes 2001). There had been a major worry
231: that the two black holes stall in their approach to each other (Valtonen
232: 1996, Yu 2003, Merritt 2005, Milosavljevi{\'{c}} \& Merritt 2003a, b, Makino
233: \& Funato 2004, Berczik et al. 2005, 2006, Matsubashi et al. 2007) before
234: they get to the emission of gravitational waves; that the loss-cone
235: mechanism for feeding stars into orbits that intersect the binary black
236: holes is too slow. However, Zier (2006) has demonstrated that direct
237: interaction with the surrounding stars slightly further outside speeds up
238: the process, and so very likely no stalling occurs. Relaxation processes due
239: to cloud/star-star interactions are rather strong, as shown by Alexander
240: (2007), using observations of our galaxy. These interactions repopulate the
241: stellar orbits in the center of the galaxy. New work by Merritt et al.
242: (2007) is consistent with Zier (2006) and Alexander (2007). Also in a series
243: of papers Sesana, Haardt, and Madau have recently shown that even in the
244: absence of two-body relaxation or gas dynamical processes, unequal mass
245: and/or eccentric binaries with the mass larger than $10^{5}$ $M_{\odot }$
246: can shrink to the gravitational wave emission regime in less than a Hubble
247: time due to the binary orbital decay by three-body interactions in the
248: gravitationally bound stellar cusps (Sesana et al. 2006, 2007a, 2007b).
249: Finally, Hayasaki (2008) has considered the "last parsec problem" under the
250: assumption of the existence of three accretion disks: one around each black
251: hole and a third one, which is circumbinary. The circumbinary disk removes
252: orbital angular momentum from the binary via the binary-disk resonant
253: interaction, however, the mass transfer to each individual black hole adds
254: orbital angular momentum to the binary. The critical parameter of the mass
255: transfer rate is such that for SMBH binaries, it becomes larger than the
256: Eddington limit, thus these binaries will merge within a Hubble time by this
257: mechanism. The angular momentum transfer from orbit to disk was already
258: considered as a key physical concept in binary stars by Biermann \& Hall
259: (1973). All these recent works suggest that by one mechanism or another the
260: SMBHs will approach each other to distances smaller than approximately one
261: hundredth pc, when the gravitational radiation becomes the dominant
262: dissipative effect. In Section 4, we analyze the characteristic timescales
263: of the dynamical friction and gravitational radiation as function of the
264: total mass, stellar distribution radius and mass ratio of the compact binary
265: and we establish the values of the transition radius and PN parameter, for
266: which the gravitational radiation is overtaking dynamical friction.
267: 
268: In Section 5, we discuss the post-Newtonian evolution of the compact
269: binaries, following Apostolatos et al. (1994) and Kidder (1995). The new
270: element is the emphasis of the role of the mass ratio as a second small
271: parameter in the formalism. The leading order conservative effect
272: contributing to the change in the orientation of spins is the spin-orbit
273: (SO) coupling. The backreaction of the gravitational radiation, which is the
274: leading order dissipative effect below the transition radius, appears at one
275: PN order higher. We show here that for the characteristic range of mass
276: ratios the spin-flip occurs during the gravitational radiation dominated
277: inspiral regime, outside ISO. In the process we evaluate the timescales for
278: the change of the spin tilt as compared to the timescales of precessional
279: motion and gravitational radiation driven inspiral. As a by-product, we are
280: able to show that for the typical mass range the so-called transitional
281: precession occurs quite rarely.
282: 
283: We interpret and discuss the resulting model in Section 6. Here, we give a
284: tentative outline of the time sequence of the activity of two merging
285: galaxies, leading to an AGN episode of the primary black hole. A recent
286: review of the generic aspects of these galaxy nuclei as sources for ultra
287: high energy cosmic rays is in (Biermann et al. 2008).
288: 
289: Finally, we summarize our findings in the concluding remarks. Following our
290: arguments about the phase just barely before the merger we propose there
291: that the superwinds in radio galaxies (Gopal-Krishna et al. 2007,
292: Gopal-Krishna \& Wiita 2006) are in this stage, as the rapidly precessing
293: jet acts just like a powerful wind. The primary goal of our paper\ is to put
294: the derived physics into observational context, so as to allow tests to be
295: done in radio and other wavelengths.
296: 
297: \section{The relevant mass ratio range}
298: 
299: In Lauer et al. (2007), the mass distribution of galactic central black
300: holes is described, confirming earlier work, and also consistent with a
301: local analysis (Roman \& Biermann 2006). Arguments based on H\"{a}ring \&
302: Rix (2004), Gott \& Turner (1977), Hickson (1982), and Press \& Schechter
303: (1974) reasoning lead to a similar result, as does a recent observational
304: survey (Ferrarese at al. 2006b). Wilson \& Colbert (1995) also find a broken
305: power law. The probability for a specific mass ratio is an integral over the
306: black hole mass distribution, folded with the rate to actually merge
307: (proportional to the capture cross-section and the relative velocity for two
308: galaxies), e.g., isomorphic to the discussion in Silk \& Takahashi (1979)
309: for the merger of clumps of different masses. The black hole mass
310: distribution $\Phi _{BH}(M_{BH})$, the number of massive central black holes
311: in galaxies per unit volume, and black hole mass interval, can be described
312: as a broken power law, from about $m_{a}\simeq 3\times 10^{6}$ $M_{\odot }$
313: to about $m_{b}\simeq 3\times 10^{9}$ $M_{\odot }$, with a break near $%
314: m_{\star }\simeq 10^{8}$ $M_{\odot }$. The lower masses have been discussed
315: in some detail by Barth et al. (2005). The values of $m_{a},~m_{b}$ and $%
316: m_{\ast }$ imply that we have two mass ranges of a factor of $30$ each. The
317: masses above $10^{8}$ $M_{\odot }$ are rapidly becoming rare with higher
318: mass, so that the lower mass range is statistically more important. That
319: ratio range is then $1:1$ to $30:1$; while in the higher mass range the
320: maximal range of the masses is also $30:1$.
321: 
322: The mass of the central massive black hole scales with the mass of the
323: spheroidal component, as with the total mass of a galaxy (the dark matter),
324: see Benson et al. (2007). The rate of black hole mergers is some fraction of
325: all mergers of massive galaxies. If, as argued by Zier (2006) the approach
326: of the two black holes does not stall, then each merger of two massive
327: galaxies will inevitably lead to the merger of the two central black holes.
328: This is supported by the statistical arguments of Rottmann (2001), using
329: radio observations, that all strong central activity in galaxies may involve
330: a merger of two black holes. Therefore, observational evidence suggests that
331: black holes do merge, and do so on the rather short timescales of AGNs.
332: 
333: The interactions and mergers of galaxies clearly depend on the three angular
334: momenta: the two intrinsic spins, and the relative orbital angular momentum,
335: as well as on the initial distance and relative velocity of the two
336: galaxies. Once all these parameters are given, the evolution is quite
337: deterministic. The observations of Gilmore et al. (2007) strongly suggest,
338: that the initial seed galaxies are today's dwarf elliptical galaxies, all of
339: which are consistent with a lower bound to a common total mass of $5\,\times
340: 10^{7}\;M_{\odot }$. This implies that all galaxies, and a fortiori all
341: central black holes, have undergone very many mergers.
342: 
343: The observations of Bouwens \& Illingworth (2006) and Iye et al. (2006)
344: strongly suggest that much of this merger history happened earlier than
345: redshift 6, perhaps mostly between redshifts $9$ and $6$. Each individual
346: merger runs along a well-defined evolutionary track, but all of these
347: mergers are completely uncorrelated with each other. Therefore, the ensemble
348: of very many mergers can be treated statistically, and this is what we
349: proceed to do, using the constant mass ratio between the spheroidal
350: component of galaxies and their central black holes. We thus use the merger
351: rate of galaxies as closely equivalent to the merger rate of the central
352: black holes.
353: 
354: The statistics of the mergers is given by the integral for the number of
355: mergers $N(q)$ per volume and time for a given mass ratio $q$, defined to be
356: larger than unity. This merger rate is the product of the distribution of
357: the first black hole with the distribution of the second black hole
358: multiplied by a rate $F$. The latter in principle depends on both the
359: cross-section and relative velocity of the two galaxies, the velocities
360: however are not very different, as the universe is not old enough for mass
361: segregation. The cross-section in turn depends on the two masses, thus $%
362: F=F(q,m)$. If we integrate for all cases, in which the first black hole is
363: less massive than the second black hole, we undercount by a factor of $2$,
364: and we have to correct for this factor. The general relationship is
365: 
366: \begin{equation}
367: N(q) = 2 \int_{m_{a}}^{m_{b} / q} \Phi_{BH} (m) \Phi_{BH} (q m) F(q, m) d m
368: \end{equation}
369: 
370: It is likely that the more massive black hole, and so the more massive host
371: galaxy, will dominate the merger rate $F$, so that it can be approximated as
372: a function of $qm$ alone, and a power law behavior with $F\sim q^{\xi }$
373: with $\xi >0$ should be adequate for a first approximation. To estimate $\xi 
374: $ roughly we just observe, that dwarf spheroidals have a core radius of a
375: few hundred pc (Gilmore et al. 2007), while our Galaxy has a core radius of
376: about $3$ kpc (Klypin et al. 2002), so a factor of $10$ in radius ($10^{2}$
377: in cross-section) for a factor of about $10^{4}$ in mass, thus the exponent
378: is likely to be approximately $1/2$; therefore a reasonable first estimate
379: for any cross-section is $\xi =1/2$. In this instance we use the approximate
380: equivalence of galaxy mergers with black hole mergers.
381: 
382: As the black hole mass distribution has a break at $q_{\ast }=30$, we use $%
383: \Phi _{BH}(m)\sim m^{-\tilde{\alpha}}$ for the first mass range, and $\Phi
384: _{BH}(m)\sim m^{-\tilde{\beta}}$ for the second. For the range $q$ from $1$
385: to $30$ we have as a dominant contribution
386: 
387: \begin{eqnarray}
388: N(q) &\sim &\int_{m_{a}}^{m_{\star }/q}\left( \frac{m}{m_{\star }}\right) ^{-%
389: \tilde{\alpha}}\left( \frac{mq}{m_{\star }}\right) ^{-\tilde{\alpha}}\left( 
390: \frac{mq}{m_{\star }}\right) ^{\xi }dm  \nonumber \\
391: &+&\int_{m_{\star }/q}^{m_{\star }}\left( \frac{m}{m_{\star }}\right) ^{-%
392: \tilde{\alpha}}\left( \frac{mq}{m_{\star }}\right) ^{-\tilde{\beta}}\left( 
393: \frac{mq}{m_{\star }}\right) ^{\xi }dm  \nonumber \\
394: &+&\int_{m_{\star }}^{m_{b}/q}\left( \frac{m}{m_{\star }}\right) ^{-\tilde{%
395: \beta}}\left( \frac{mq}{m_{\star }}\right) ^{-\tilde{\beta}}\left( \frac{mq}{%
396: m_{\star }}\right) ^{\xi }dm
397: \end{eqnarray}%
398: and for the case of $q$ above $30$ we have the contribution
399: 
400: \begin{equation}
401: N(q)\sim \int_{m_{a}}^{m_{b}/q}\left( \frac{m}{m_{\star }}\right) ^{-\tilde{%
402: \alpha}}\left( \frac{mq}{m_{\star }}\right) ^{-\tilde{\beta}}\left( \frac{mq%
403: }{m_{\star }}\right) ^{\xi }dm~.
404: \end{equation}
405: 
406: The various models shown in Lauer et al.(2007) show that a range of values
407: of $\tilde{\alpha}$ and $\tilde{\beta}$ is possible, with $\tilde{\alpha}$
408: ranging between approximately $1$ and $2$, and $\tilde{\beta}$ from $3$ to
409: larger values. Benson et al.(2007) propose $\tilde{\alpha}\approx 0.65$. We
410: adopt here the approximate values for $\tilde{\alpha}$ and $\tilde{\beta}$
411: of $1$ and $3$, to be cautious, and for $\xi $ we adopt $1/2$. With these
412: values the above integrands are monotonically decreasing functions and the
413: integrals are dominated by the lower limits. Thus, the four terms scale with 
414: $q$ as ${q}^{\xi -\tilde{\alpha}}$, ${q}^{-1+\tilde{\alpha}}$, ${q}^{\xi -%
415: \tilde{\beta}}$, and again ${q}^{\xi -\tilde{\beta}}$.
416: 
417: Let us consider the four terms: the first term is small galaxies merging
418: with small galaxies, and so not very interesting, as the cross-section is
419: low. However, for this distribution the number of mergers in the mass ratio
420: range $30:1$ to $3:1$ versus $3:1$ to $1:1$ is about $5$. The more extreme
421: mass ratios are more common. For the second term this ratio of mergers in
422: the two mass ratio ranges is about $14$. As this is massive galaxies merging
423: with smaller galaxies (above and below the break $m_{\star }$), this is the
424: most interesting case, and also quite common. The third term is almost
425: negligible, and the fourth term adds cases to the second term with more
426: extreme mass ratios, beyond $30:1$, and so emphasizes the large mass ratio
427: range.
428: 
429: So, among the relevant cases the rate of mergers of mass ratio of more than $%
430: 3:1$ to those with a smaller mass ratio is in the range of $5:1$ to $14:1$,
431: about an order of magnitude. Focussing on those cases where one black hole
432: is at $10^{8}$ $M_{\odot }$ or larger, the ratio is larger than $14:1$.
433: Speculating that the exponent $\xi $ could be larger would enhance all these
434: effects; enlarging $\tilde{\alpha}$ would weaken them. Therefore we will
435: deal in the following with this much more common extended mass ratio range $%
436: 30:1$ to $3:1$, which as will be shown, allows to use analytical methods.
437: 
438: \section{The spin and orbital angular momentum in the PN regime}
439: 
440: We assume the compact binary system to be composed of two masses $m_{i}$, $%
441: i=1,2$ , each having the spin $\mathbf{S}_{\mathbf{i}}$. By definition, the
442: characteristic radius $R_{i}$ of compact objects is of \ the same order of
443: magnitude that the gravitational radius $R_{G}=Gm_{i}/c^{2}$ (where $c$ is
444: the velocity of light and $G$ is the gravitational constant). Therefore, the
445: magnitude of the spin vector can be approximated as $S_{i}\approx
446: m_{i}R_{i}V_{i}\approx Gm_{i}^{2}V_{i}/c^{2}$, where $V_{i}$ is the
447: characteristic rotation velocity of the $i$th compact object \ As black
448: holes rotate fast due to accretion, $V_{i}/c$ is of order unity.
449: Equivalently we can introduce $S_{i}=(G/c)m_{i}^{2}\chi _{i}$, with $\chi
450: _{i}$ being the dimensionless spin parameter. Then maximal rotation implies $%
451: \chi _{i}=1$.
452: 
453: The PN expansion is done in terms of the small parameter 
454: \begin{equation}
455: \varepsilon \approx \frac{Gm}{c^{2}r}\approx \frac{v^{2}}{c^{2}},
456: \end{equation}%
457: where $m=m_{1}+m_{2}$ is the total mass and $v$ is the orbital velocity of
458: the reduced mass particle $\mu =m_{1}m_{2}/m$, which is in orbit about the
459: fixed mass $m$ (according to the one-centre problem in celestial mechanics).
460: The two expressions for $\varepsilon $ are of the same order of magnitude
461: due to the virial theorem. As in certain expressions odd powers of $v/c$ may
462: occur, it is common to have half-integer orders in the post-Newtonian
463: treatment of the inspiral of a compact binary system.
464: 
465: Whenever the masses of the two compact objects are comparable, either of $%
466: Gm_{i}/c^{2}r$ also represent one post-Newtonian order. However, as we have
467: argued before, for colliding galactic black holes it is typical that their
468: masses differ by 1 order of magnitude, so that we have a second small
469: parameter in the formalism. By choosing $m_{2}$ as the smaller mass, we can
470: also define the mass ratio 
471: \begin{equation}
472: \nu =\frac{m_{2}}{m_{1}}=\frac{1}{q}\in \left( 0,1\right) ~.
473: \end{equation}%
474: In the literature the symmetric mass ratio 
475: \begin{equation}
476: \eta =\frac{\mu }{m}\in \left( 0,\frac{1}{4}\right) ~
477: \end{equation}%
478: is also frequently employed. The two mass ratios are related as%
479: \begin{equation}
480: \eta =\frac{\nu }{\left( 1+\nu \right) ^{2}}~,
481: \end{equation}%
482: and for small $\nu $ we have $\eta =\allowbreak \nu -2\nu ^{2}+O\left( \nu
483: ^{3}\right) $.
484: 
485: For the typical mass ratio range of SMBH binaries either $\eta $ or $\nu $
486: can be chosen as the second small parameter in the formalism. However, while
487: these stay constant, the PN parameter $\varepsilon $ evolves during the
488: inspiral toward higher values. Indeed, the separation of the components of
489: the binary with $m=10^{8}M_{\odot }$ evolves as 
490: \begin{equation}
491: r=\frac{Gm}{c^{2}\varepsilon }=\frac{r_{S}}{2\varepsilon }=\allowbreak
492: 4.\,\allowbreak 781\,3\times 10^{-6}\frac{{pc}}{\varepsilon }{,}  \label{rep}
493: \end{equation}%
494: where $r_{S}$ represents the Schwarzschild radius. The interaction of the
495: galactic black holes with the stellar environment begins when the black
496: holes are a few kpc away from each other (then $\varepsilon \approx 10^{-8}$%
497: ). The dynamical friction becomes subdominant at about $0.005$ pc (Zier \&
498: Biermann 2001, Zier 2006), when the gravitational radiation becomes the
499: leading dissipative effect. Thus, $\varepsilon =\varepsilon ^{\ast }\approx
500: 10^{-3}$ is the value of the PN parameter for which the gravitational
501: radiation is driving the dissipation of energy and orbital angular momentum.
502: Then follows the inspiral stage of the evolution of compact binaries, which
503: continues until the domain of validity of the post-Newtonian approach is
504: reached, at few gravitational radii, at ISO. Further away a numerical
505: treatment is necessary in order to describe the plunge, which is finally
506: followed by the ring-down. The PN formalism can be considered valid until $%
507: \varepsilon \approx 10^{-1}$.
508: 
509: Theoretically, it is possible for a small $\nu $, that at certain stage of
510: the inspiral, the increasing $\varepsilon ^{1/2}$ becomes of the same order
511: of magnitude as $\nu $ and later on it even exceeds $\nu $. Such a situation
512: would shift the numerical value of several contributions to various physical
513: quantities into the range of higher or lower PN orders, depending of the
514: involved power of $\nu $. 
515: \begin{table*}[tbp]
516: \centering%
517: \begin{minipage}{140mm}
518: \caption{The evolution of the ratio 
519: $S_{1}/L\approx \protect\varepsilon^{1/2}\protect\nu ^{-1}$ 
520: in the range 
521: $\protect\varepsilon =10^{-3}\div 10^{-1}$ 
522: for various values of the mass ratio $\protect\nu $. }
523: \label{Table1}
524: \begin{center}
525: \begin{tabular}{lrlrl}
526: $S_{1}/L=\varepsilon^{1/2}\nu ^{-1}$ & 
527: 					& 
528: $\varepsilon \approx 10^{-3}$ &  
529: 					& 
530: $\varepsilon \approx 10^{-1}$ \\
531: \hline\hline
532: $\nu =1$ & $0.03$ & $\left( S_{1}\ll L\right) $ &
533:  \qquad $0.3$ & $\left( S_{1}< L\right) $ \\
534: $\nu =1/3$ & $0.1$ & $\left( S_{1}< L\right) $ & 
535:  \qquad $1$ & $\left(S_{1}\approx L\right) $ \\
536: $\nu =1/30$ & $1$ & $\left( S_{1}\approx L\right) $ & 
537:  \qquad $10$ & $\left(S_{1}> L\right) $ \\
538: $\nu =1/900$ & $30$ & $\left( S_{1}\gg L\right) $ & 
539:  \qquad $300$ & $\left(S_{1}\gg L\right) $
540: \end{tabular}
541: \end{center}
542: \end{minipage}
543: \end{table*}
544: 
545: The spin ratio (for similar rotation velocities $V_{1}\approx V_{2}$) can be
546: expressed as%
547: \begin{equation}
548: \frac{S_{2}}{S_{1}}\approx \left( \frac{m_{2}}{m_{1}}\right) ^{2}=\nu ^{2}{~.%
549: }  \label{S1S2}
550: \end{equation}%
551: The ratio of the spins to the orbital angular momentum becomes%
552: \begin{eqnarray}
553: \frac{S_{2}}{L} &\approx &\frac{Gm_{2}^{2}V_{2}/c^{2}}{\mu rv}=\left( \frac{%
554: Gm}{c^{2}r}\right) \left( \frac{c}{v}\right) \left( \frac{V_{2}}{c}\right) 
555: \frac{m_{2}}{m_{1}}\approx \varepsilon ^{1/2}\nu ~, \\
556: \frac{S_{1}}{L} &=&\frac{S_{2}}{L}\frac{S_{1}}{S_{2}}\approx \varepsilon
557: ^{1/2}\nu ^{-1}~{.}  \label{S1L}
558: \end{eqnarray}%
559: We note that the approximations in the above formulae (\ref{S1S2})-(\ref{S1L}%
560: ) are related only to the fact that we have assumed maximal rotation (thus $%
561: V_{i}/c\lesssim 1$). First, we note that the above ratios involving the
562: spins of the compact objects already contain $\varepsilon ^{1/2}$. Thus, the
563: counting of the inverse powers of $c^{2}$ is not equivalent with the PN
564: order, when compact objects are involved. Further, while the ratio $S_{2}/L$
565: is shifted toward higher orders by a small $\nu $ (therefore $S_{2}\ll L$
566: during all stages of the inspiral), the order of the ratio of the spin of
567: the dominant black hole to the magnitude of the orbital angular momentum is
568: not fixed. Indeed, it is determined by the relative magnitude of the small
569: parameters $\varepsilon $ and $\nu $. As $\varepsilon $ increases during the
570: inspiral, whenever $\nu $ falls in the range of $\varepsilon ^{1/2}$, the
571: initial epoch with $S_{1}<L$ is followed by $S_{1}\approx L$ and $S_{1}>L$
572: epochs (Table \ref{Table1}).
573: 
574: We have concluded in the previous section that the range of mass ratios $q$
575: between $3:1$ and $30:1$ is the most common. For such binaries\ the sequence
576: of the three epochs $S_{1}<L$, $S_{1}\approx L$, and $S_{1}>L$ is fairly
577: representative. We call this \textit{intermediate mass ratio mergers}, which
578: has to be contrasted with the case of equal mass mergers, where the orbital
579: angular momentum dominates throughout the inspiral; and with the case of
580: extreme mass ratio mergers (which we define as having mass ratios larger
581: than $30:1$), where the larger spin dominates from the beginning of the
582: inspiral to the end of the PN phase, as can be seen from our Table \ref%
583: {Table1}.
584: 
585: \section{The timescales}
586: 
587: The value $\varepsilon \approx 10^{-3}$ from which the PN analysis with the
588: gravitational radiation as the leading dissipative effect can be applied was
589: adopted in the previous section for a compact binary with total mass $%
590: m=10^{8}\,M_{\odot }$ and mass ratio $\nu =10^{-1}$. This was based on the
591: analysis in Zier \& Biermann (2001) and Zier (2006), where it was shown that
592: at around $5\times 10^{-3}$ pc gravitational radiation takes over from
593: dynamical friction in the interaction with stars in the angular momentum
594: loss of the black hole binary. Further arguments for the binary to reach the
595: gravitational wave emission regime were presented by Alexander (2007),
596: Sesana et al. (2006, 2007a, 2007b), and Hayasaki (2008).
597: 
598: In this section we raise the question, whether the value of the transition
599: radius (and the corresponding value of the PN parameter) depends on $m$ and $%
600: \nu $. In order to answer this, we compare the characteristic timescales of
601: the gravitational radiation and dynamical friction.
602: 
603: The timescale of gravitational radiation (as will be derived in Section 5)
604: is 
605: \begin{equation}
606: \frac{1}{t_{gw}}=-\frac{\dot{L}}{L}\approx \frac{32c^{3}}{5Gm}\varepsilon
607: ^{4}\eta ~.
608: \end{equation}%
609: The timescale for the secondary black hole to lose angular momentum by
610: gravitational interaction with the surrounding stellar distribution is
611: (Binney \& Tremaine 1987) 
612: \begin{equation}
613: t_{fr}=\frac{v^{3}}{2\pi G^{2}m_{2}\rho _{distr}\Lambda }\left( \frac{\Delta
614: v}{v}\right) ^{2}~.
615: \end{equation}%
616: With the maximally allowable change in the velocity $\Delta v/v=1$ we find
617: the relevant \textit{full} timescale. Here, $\Lambda =\ln (b_{max}/b_{min})$
618: is the logarithm of the ratio of the maximal distance within the system,
619: divided by the typical distance between objects. The latter is large for
620: clouds, so the ratio is low and $\Lambda $ of the order unity, while for
621: stars or dark matter particles $\Lambda $ can be taken as $10\div 20$. For
622: the merger, the estimate based on clouds is more appropriate, thus,
623: following the reasoning of Binney \& Tremaine (1987) we adopt $\Lambda =3$.
624: The compact stellar distribution with density $\rho _{distr}$, radius $%
625: r_{distr}$, and mass $m_{distr}$ is of the same order in mass as the black
626: hole binary (Zier \& Biermann 2001, Ferrarese et al. 2006a) with a scale $%
627: r_{distr}$ of a few pc, and so under the assumption of a spherically
628: symmetric distribution we set%
629: \begin{equation}
630: \rho _{distr}=\frac{3m}{4\pi r_{distr}^{3}}~.
631: \end{equation}%
632: By employing the definitions of the PN parameter and mass ratio, we get%
633: \begin{equation}
634: \frac{1}{t_{fr}}=\frac{9G^{2}}{2c^{3}}\frac{m^{2}}{r_{distr}^{3}}\varepsilon
635: ^{-3/2}\eta \left( 1+\nu \right) ~.
636: \end{equation}%
637: This gives the full timescale of the dynamical friction.
638: 
639: The two timescales become comparable for a PN parameter:%
640: \begin{equation}
641: \varepsilon ^{\ast }=K\left( \nu \right) \left( \frac{Gm}{c^{2}r_{distr}}%
642: \right) ^{6/11}~,
643: \end{equation}%
644: with $K\left( \nu \right) $ being a factor of order unity, defined as:%
645: \begin{equation}
646: K\left( \nu \right) =\left[ \frac{45}{64}\left( 1+\nu \right) \right]
647: ^{2/11}\in \left( \allowbreak 0.938,~1.\,\allowbreak 064\right) ~.
648: \end{equation}%
649: corresponding to the distance $r^{\ast }$%
650: \begin{equation}
651: r^{\ast }=K^{-1}\left( \nu \right) \left( \frac{Gm}{c^{2}}\right)
652: ^{5/11}r_{distr}^{6/11}~.
653: \end{equation}%
654: Notably, the dependence on the mass ratio is rather weak and in practice it
655: can be neglected. Inserting then for $r_{distr}=5$ pc and using as a
656: reference value for the mass $m=10^{8}~M_{\odot }$, we obtain $\varepsilon
657: ^{\ast }\approx 10^{-3}$ and $r^{\ast }\approx 0.005$ pc in agreement with
658: the discussion in Zier \& Biermann (2001). We also note that in fact any
659: other reasonable value for $\Lambda $ and $\Delta v/v$ will give a factor of
660: order unity in Eq. (14), as this number arises by taking the power $2/11$.
661: The weak dependence on $\nu $ through $K$ is due to the same reason.
662: 
663: The value of the PN parameter thus scales with $m^{6/11}$ and radius $%
664: r_{distr}^{-6/11}$ as%
665: \begin{equation}
666: \varepsilon ^{\ast }\approx 10^{-3}\left( \frac{m}{10^{8}~M_{\odot }}\right)
667: ^{6/11}\left( \frac{5{\ pc}}{r_{distr}}\right) ^{6/11}~,
668: \end{equation}%
669: while the transition radius scales with $m^{5/11}$ and $r_{distr}^{6/11}$ as 
670: \begin{equation}
671: r^{\ast }\approx 0.005~pc~\left( \frac{m}{10^{8}M_{\odot }}\right)
672: ^{5/11}\left( \frac{r_{distr}}{5\ {\ }pc}\right) ^{6/11}~.
673: \end{equation}%
674: For $m=10^{9}M_{\odot }$ and for the same radius of stellar distribution
675: then $r^{\ast }\approx $ $0.01$ pc.
676: 
677: We conclude that both the dependence of the transition radius and the
678: corresponding PN parameter on the total mass and on the stellar distribution
679: radius are weak, while there is practically no dependence on the mass ratio.
680: 
681: \section{The inspiral of spinning compact binaries in the gravitational
682: radiation dominated regime}
683: 
684: In the first subsection of this section we present the conservative dynamics
685: of an isolated compact binary in a post-Newtonian treatment, emphasizing the
686: role of the second small parameter, as a new element. Then in the second
687: subsection we take into account the effect of gravitational radiation,
688: deriving how the spin-flip occurs for the typical mass ratio range. The
689: limits of validity of our results obtained by using these two small
690: parameters will be considered below in subsection 5.3.
691: 
692: \subsection{Conservative dynamics below the transition radius}
693: 
694: The interchange in the dominance of either $L$ or $S_{1}$ has a drastic
695: consequence on the dynamics of the compact binary. To see this, let us
696: summarize first the conservative dynamics, valid up to the second
697: post-Newtonian order. The constants of the motion are the total energy $E$
698: and the total angular momentum vector $\mathbf{J=L+S}_{\mathbf{1}}+\mathbf{S}%
699: _{\mathbf{2}}$ (Kidder et al. 1993). The angular momenta $\mathbf{L,~S}_{%
700: \mathbf{i}}$ are not conserved separately. The spins obey a precessional
701: motion (Barker \& O'Connell 1975, Barker \& O'Connell 1979):%
702: \begin{equation}
703: \mathbf{\dot{S}_{i}}=\mathbf{\Omega }_{\mathbf{i}}\times \mathbf{S_{i}}\ ,
704: \label{spinprec}
705: \end{equation}%
706: with the angular velocities given as a sum of the spin-orbit, spin-spin, and
707: quadrupole-monopole contributions. The latter come from regarding one of the
708: binary components as a mass monopole moving in the quadrupolar field of the
709: other component.
710: 
711: The leading order contribution due to the SO interaction (discussed in
712: Kidder et al. 1993, Apostolatos et al. 1994, Kidder 1995, Ryan 1996, Rieth
713: \& Sch\"{a}fer 1997, Gergely at al. 1998a, 1998b, 1998c, O'Connell 2004),
714: cause the spin axes to tumble and precess. The spin-spin (Kidder 1995,
715: Apostolatos 1995, Apostolatos 1996, Gergely 2000a, 2000b), mass quadrupolar
716: (Poisson 1998, Gergely \& Keresztes 2003, Flanagan \& Hinderer 2007, Racine
717: 2008), magnetic dipolar (Ioka \& Taniguchi 2000, Vas\'{u}th et al. 2003),
718: self-spin (Mik\'{o}czi et al. 2005) and higher order spin-orbit effects
719: (Faye et al. 2006, Blanchet et al. 2006) slightly modulate this process.
720: 
721: The SO precession occurs with the angular velocities 
722: \begin{eqnarray}
723: \mathbf{\Omega }_{\mathbf{1}} &=&{\frac{G\left( 4+3\nu \right) }{2c^{2}r^{3}}%
724: }\mathbf{L_{N}}~,  \label{Om1} \\
725: \mathbf{\Omega }_{\mathbf{2}} &=&{\frac{G\left( 4+3\nu ^{-1}\right) }{%
726: 2c^{2}r^{3}}}\mathbf{L_{N}}~,  \label{Om2}
727: \end{eqnarray}%
728: where $\mathbf{L_{N}}=\mu \mathbf{r}\times \mathbf{v}$ is the Newtonian part
729: of the orbital angular momentum. The total orbital angular momentum $\mathbf{%
730: L}$ also contains a contribution $\mathbf{L}_{\mathbf{SO}}$ (Kidder 1995),
731: which for compact binaries is of the order of $\varepsilon ^{3/2}L_{N}$ .
732: 
733: Due to the conservation of $\mathbf{J}$, the orbital angular momentum
734: evolves as
735: 
736: \begin{equation}
737: \mathbf{\dot{L}=}{\frac{G}{2c^{2}r^{3}}}\left[ \left( 4+3\nu \right) \mathbf{%
738: S_{1}}+{\left( 4+3\nu ^{-1}\right) }\mathbf{S_{2}}\right] \mathbf{\times L~.}
739: \end{equation}%
740: (By adding a correction term of order $\varepsilon ^{3/2}$ relative to the
741: leading order terms,$\ $we have changed $\mathbf{L_{N}}$ into $\mathbf{L}$
742: on the right-hand side of the above equation.) \ 
743: 
744: To leading order in $\nu $ we obtain:%
745: \begin{eqnarray}
746: \mathbf{\dot{S}_{1}} &=&{\frac{2G}{c^{2}r^{3}}}\mathbf{L}\times \mathbf{S_{1}%
747: }\ ,  \label{Sdot} \\
748: \mathbf{\dot{L}} &\mathbf{=}&{\frac{2G}{c^{2}r^{3}}}\mathbf{S_{1}\times L~.}
749: \label{Ldotconserv}
750: \end{eqnarray}%
751: (Again, an $\mathbf{L}_{\mathbf{SO}}$ term was added to $\mathbf{L_{N}}$ on
752: the right-hand side of Eq. (\ref{Sdot}), in order to have a pure precession
753: of $\mathbf{S_{1}}$.)
754: 
755: Thus, the leading order conservative dynamics gives the following picture:
756: the dominant spin $\mathbf{S_{1}}$ undergoes a pure precession about $%
757: \mathbf{L}$, while$\ \mathbf{L}$ does the same about $\mathbf{S}_{\mathbf{1}%
758: } $. Despite the precession (\ref{Om2}), the spin $\mathbf{S}_{\mathbf{2}}$
759: can be ignored to leading order, as its magnitude is $\nu ^{2}$ times
760: smaller than $S_{1}$, e.g., Eq. (\ref{S1S2}). By adding the vanishing terms $%
761: \left( {2G/}c^{2}r^{3}\right) \mathbf{S}_{\mathbf{1}}\times \mathbf{S_{1}}$
762: and $\left( {2G/}c^{2}r^{3}\right) \mathbf{L}\times \mathbf{L}$ to the
763: right-hand sides of Eqs. (\ref{Sdot}) and (\ref{Ldotconserv}), respectively,
764: we obtain%
765: \begin{eqnarray}
766: \mathbf{\dot{S}_{1}} &=&{\frac{2G}{c^{2}r^{3}}}\mathbf{J}\times \mathbf{S_{1}%
767: }\ ,  \label{SdotJ} \\
768: \mathbf{\dot{L}} &\mathbf{=}&{\frac{2G}{c^{2}r^{3}}}\mathbf{J\times L~.}
769: \label{LdotJcons}
770: \end{eqnarray}%
771: Thus, the precessions can also be imagined to happen about $\mathbf{J}$,
772: which represents an invariant direction in the conservative dynamics up to
773: 2PN.
774: 
775: Higher order contributions to the conservative dynamics slightly modulate
776: this precessional motion. In fact, for both the spin-spin and
777: quadrupole-monopole perturbations an angular average $\bar{L}$ can be
778: introduced, which is conserved up to the 2PN order (Gergely 2000a). As $\bar{%
779: L}$ differs from $L$ just by terms of order 2PN, and $\bar{L}$ is conserved,
780: the real evolution of $\mathbf{L}$ differs from a pure precession only
781: slightly.
782: 
783: Finally, we note that as the SO precessions are 1.5 PN effects and the
784: gravitational radiation appears at 2.5 PN, at the transition radius the SO
785: precession timescale is $\varepsilon ^{-1}$ times shorter than the timescale
786: of dynamical friction. The modifications induced by the precessions in the
787: angular momentum transfer toward the stellar environment will be discussed
788: elsewhere (Zier et al. 2009, in preparation).
789: 
790: \subsection{Dissipative dynamics below the transition radius}
791: 
792: \begin{figure*}[tbp]
793: \centering\includegraphics[height=8cm]{f1.eps}
794: \caption{The old jet points in the direction of the original spin $\mathbf{%
795: S_{1}}$. When the two black holes approach each other due to the motion of
796: their host galaxies, a slow precessional motion of both the spin and of the
797: orbital angular momentum $\mathbf{L}$ begins (left figure) about the
798: direction of the total angular momentum $\mathbf{J}$, which is due to the
799: spin-orbit interaction. Gravitational radiation carries away both energy and
800: angular momentum from the system, such that the direction of $\mathbf{J}$
801: stays unchanged. As a consequence the precessional orbit slowly shrinks and
802: the magnitude of $\mathbf{L}$ decreases. This is accompanied by a continuous
803: increase in the angle $\protect\alpha $ and a decrease in $\protect\beta $.
804: When the magnitudes of $\mathbf{L}$ and $\mathbf{S}$ become comparable
805: (middle figure), the precessional motions are much faster (for typical
806: values see Table \protect\ref{Table2}). In the typical mass ratio range $%
807: \protect\nu =1/3\div 1/30$ the magnitude of $\mathbf{L}$ becomes small as
808: compared to the magnitude of the spin, which is unchanged by gravitational
809: radiation (except the upper boundary of the range at $\protect\nu =1/3$ when 
810: $L$ and $S$\ are still comparable). Before reaching the innermost stable
811: orbit, the spin becomes almost aligned to the (original direction of the)
812: total angular momentum, and a new jet can form along this direction.
813: Therefore, for the typical mass ratio range the spin-flip phenomenon has
814: occurred in the inspiral phase and not much orbital angular momentum is left
815: to modify the direction of the spin during the plunge and ring-down. In the
816: regime in between the initial and final states the precessing jet acts as a
817: superwind, sweeping away the environment of the jets.}
818: \label{Fig1}
819: \end{figure*}
820: 
821: Dynamics becomes dissipative from 2.5 PN orders. Then gravitational
822: quadrupolar radiation carries away both energy and angular momentum. Orbital
823: eccentricity is dissipated faster, than the rate of orbital inspiral (Peters
824: 1964), thus the orbit will circularize.\footnote{%
825: As we focus here on spin-flips, we will not dwell on possible recoil as a
826: result of the momentum loss due to gravitational radiation in the merger of
827: the two black holes (see for example Br{\"{u}}gmann 2008, Gonzalez et al.
828: 2007a, b). The accuracy of determining the distance between two separate and
829: independent active black holes (Marcaide \& Shapiro 1983, Brunthaler et al.
830: 2005) is reaching a precision, which may soon allow for recoil to be
831: measurable; no such evidence has been detected yet.} Considering circular
832: orbits and averaging over one orbit gives the following dissipative change
833: in $\mathbf{L}$:%
834: \begin{equation}
835: \mathbf{\dot{L}}^{gw}=-\frac{32G\mu ^{2}}{5r}\left( \frac{Gm}{c^{2}r}\right)
836: ^{5/2}\mathbf{\hat{L}~,}  \label{Ldotdis}
837: \end{equation}%
838: where $\mathbf{\hat{L}}$ represents the direction of the orbital angular
839: momentum. Then the total change in $\mathbf{L}$ is given by the sum of Eqs. (%
840: \ref{Ldotconserv}) and (\ref{Ldotdis}).
841: 
842: The spin-induced quasi-precessional effects both modulate the dynamics and
843: they have an important effect on gravitational wave detection (see Lang \&
844: Hughes 2006, 2008, Racine 2008, Gergely \& Mik\'{o}czi 2008).
845: 
846: The dissipative dynamics, with the inclusion of the leading order SO
847: precessions and the dissipative effects due to gravitational radiation,
848: averaged over circular orbits, was discussed in detail in Apostolatos et al.
849: (1994) for the one-spin case $\mathbf{S_{2}}=0$ and equal masses $\nu =1$.
850: For the typical mass ratio $\nu \in \left( 1/30,~1/3\right) $, and keeping
851: only the leading order contributions in the $\nu $-expansion also gives $%
852: \mathbf{S_{2}}=0$ (the leading order contributions in $\mathbf{S_{2}}$ are
853: of order $\nu ^{2}$). In this subsection we will analyze in depth the
854: angular evolutions and the timescales involved.
855: 
856: As for any vector $\mathbf{X}$ with magnitude $X$ and direction $\mathbf{%
857: \hat{X}}$ one has $\mathbf{\dot{X}}=X\mathbf{\skew{0}{\dot}{\hat{X}}}+\dot{X}%
858: \mathbf{\hat{X}}$, the change in the direction can be expressed as $\mathbf{%
859: \skew{0}{\dot}{\hat{X}}}=\left( \mathbf{\dot{X}}-\dot{X}\mathbf{\hat{X}}%
860: \right) /X$. Also the identity $X^{2}=\mathbf{X}^{2}$ gives $\dot{X}=\mathbf{%
861: \hat{X}}\cdot \mathbf{\dot{X}}$. Then Eqs. (\ref{SdotJ})-(\ref{Ldotdis})
862: imply%
863: \begin{eqnarray}
864: \dot{S}_{1} &=&0~,  \nonumber \\
865: \mathbf{\skew{0}{\dot}{\hat{S}}_{1}} &=&{\frac{2G}{c^{2}r^{3}}}\mathbf{J}%
866: \times \mathbf{\hat{S}_{1}}\ ,  \nonumber \\
867: \dot{L} &=&-\frac{32G\mu ^{2}}{5r}\left( \frac{Gm}{c^{2}r}\right) ^{5/2}~, 
868: \nonumber \\
869: \mathbf{\skew{0}{\dot}{\hat{L}}} &=&{\frac{2G}{c^{2}r^{3}}}\mathbf{J\times 
870: \hat{L}}~.  \label{dynamics}
871: \end{eqnarray}%
872: The total angular momentum $\mathbf{J}$ is also changed by the emitted
873: gravitational radiation. As no other change occurs up the 2PN orders, $%
874: \mathbf{\dot{J}=}\dot{L}\mathbf{\hat{L}}$ and%
875: \begin{eqnarray}
876: \dot{J} &=&\dot{L}\left( \mathbf{\hat{L}}\cdot \mathbf{\hat{J}}\right) 
877: \mathbf{~,}  \nonumber \\
878: \mathbf{\skew{0}{\dot}{\hat{J}}} &=&\frac{\dot{L}}{J}\left[ \mathbf{\hat{L}}%
879: -\left( \mathbf{\hat{L}}\cdot \mathbf{\hat{J}}\right) \mathbf{\hat{J}}\right]
880: ~.  \label{Jdot}
881: \end{eqnarray}%
882: Note that from the second Eq. (\ref{Jdot}) it is immediate that the
883: direction of $\mathbf{J}$ changes violently, whenever $J$ is small compared
884: to $\dot{L}$.
885: 
886: To leading order in $\nu $ , the vectors $\mathbf{L}$, $\mathbf{S}_{\mathbf{1%
887: }}$, and $\mathbf{J}$ form a parallelogram, characterized by the angles $%
888: \alpha =\cos ^{-1}\left( \mathbf{\hat{L}\cdot \hat{J}}\right) $ and $\beta
889: =\cos ^{-1}\left( \mathbf{\hat{S}}_{\mathbf{1}}\mathbf{\cdot \hat{J}}\right) 
890: $. From Eqs.(\ref{dynamics}) and (\ref{Jdot}), we obtain%
891: \begin{eqnarray}
892: \dot{\alpha} &=&-\frac{\dot{L}}{J}\sin \alpha >0~,  \label{alphadot} \\
893: \dot{\beta} &=&\frac{\dot{L}}{J}\sin \alpha <0~.  \label{betadot}
894: \end{eqnarray}%
895: In the latter equation, we have used that $\mathbf{\hat{S}}_{\mathbf{1}}%
896: \mathbf{\cdot \hat{L}=\cos }\left( \alpha +\beta \right) $.
897: 
898: Thus, we have found the following picture for the inspiral of the compact
899: binary after the transition radius. By disregarding gravitational radiation,
900: the SO precessions (\ref{Sdot}) and (\ref{Ldotconserv}) assure that the
901: vectors $\mathbf{L}$ and $\mathbf{S_{1}}$ are precessing about $\mathbf{J}$
902: (a fixed direction), but also about each other (then the respective axes of
903: precession evolve in time). Gravitational radiation slightly perturbs this
904: picture. The angle $\alpha +\beta $ between the orbital angular momentum and
905: the dominant spin stays constant during the inspiral, even with the
906: gravitational radiation taken into account. By contrast, the angle between $%
907: \mathbf{J}$ and$\ \mathbf{L}$ continuously increases, while the angle between%
908: $\ \mathbf{J}$ and $\mathbf{S_{1}}$ decreases with the same rate. This also
909: means that due to gravitational radiation, the vectors $\mathbf{L}$ and $%
910: \mathbf{S_{1}}$ do not precess about $\mathbf{J}$ any more in an exact
911: sense. They keep precessing about each other, however.
912: 
913: The change in the total angular momentum $\mathbf{\dot{J}=}\dot{L}\mathbf{%
914: \hat{L}}$ is about the orbital angular momentum, which in turn basically
915: (disregarding gravitational radiation) undergoes a precessional motion about 
916: $\mathbf{J.}$ This shows that the averaged change in $\mathbf{J}$ is along $%
917: \mathbf{J}$. This conclusion, however, depends strongly on whether the
918: precessional angular frequency $\Omega _{p}$ is much higher than the change
919: in the angles $\alpha $ and $\beta $. Indeed, if these are comparable, the
920: component perpendicular to $\mathbf{J}$ in the change $\mathbf{\dot{J}=}\dot{%
921: L}\mathbf{\hat{L}}$ will not average out during one precessional cycle, as
922: due to the increase of $\alpha $ it can significantly differ at the
923: beginning and at the end of the same precessional cycle, see Fig \ref{Fig1}.
924: 
925: The regime with $\Omega _{p}\gg \dot{\alpha}$ can be well approximated by a
926: precessional motion of both $\mathbf{L}$ and $\mathbf{S_{1}}$ about a fixed $%
927: \mathbf{\hat{J}}$, with the magnitudes of $\mathbf{L}$ and $\mathbf{J}$
928: slowly shrinking, the angle $\alpha $ slowly increasing and $\beta $ slowly
929: decreasing. As a result, during the inspiral, the orbital angular momentum
930: slowly turns away from $\mathbf{J}$, while $\mathbf{S_{1}}$ slowly
931: approaches the direction of $\mathbf{J}$. This regime is characteristic for
932: the majority of cases, and it was called simple precession in Apostolatos et
933: al. (1994).
934: 
935: Whenever $\Omega _{p}\approx \dot{\alpha}$, the conclusion of having $%
936: \mathbf{\dot{J}}$ in the direction of $\mathbf{\hat{J}}$ does not hold for
937: the average over one precession. This results in a change in the direction
938: of $\mathbf{J}$ in each precessional cycle. The evolution becomes much more
939: complicated (in fact no approximate analytical solution is known), and it
940: was called transitional precession in Apostolatos et al. (1994).
941: 
942: Let us see now when the two types of evolution typically occur. For this, we
943: note that the inspiral rate $\dot{L}/L$ is of the order 
944: \begin{equation}
945: \frac{\dot{L}}{L}\approx -\frac{32c^{3}}{5Gm}\varepsilon ^{4}\eta ~,
946: \end{equation}%
947: while the precessional angular velocity $\Omega _{p}=2GJ/c^{2}r^{3}$ gives
948: the estimate%
949: \begin{equation}
950: \Omega _{p}\mathbf{\approx }\frac{2c^{3}}{Gm}{\varepsilon }^{5/2}\eta \frac{{%
951: J}}{L}~.  \label{Omegapepeta}
952: \end{equation}%
953: Finally, the tilt velocity of $\mathbf{L}$ is of the order%
954: \begin{eqnarray}
955: \dot{\alpha} &\approx &\frac{32c^{3}}{5Gm}\varepsilon ^{7/2}\eta \nu \left( 
956: \frac{S_{1}}{J}\right) ^{2}\sin \left( \alpha +\beta \right)  \nonumber \\
957: &\approx &\frac{32c^{3}}{5Gm}\varepsilon ^{9/2}\eta \nu ^{-1}\left( \frac{L}{%
958: J}\right) ^{2}\sin \left( \alpha +\beta \right)  \label{alphadotepeta}
959: \end{eqnarray}
960: We have used 
961: \begin{equation}
962: \sin \alpha =\frac{S_{1}}{J}\sin \left( \alpha +\beta \right) \approx
963: \varepsilon ^{1/2}\nu ^{-1}\frac{L}{J}\sin \left( \alpha +\beta \right)
964: \label{albe}
965: \end{equation}
966: for the first and second expressions of $\dot{\alpha}$ , respectively,
967: together with Eq. (\ref{S1L}). For comparison, these are all represented in
968: Table \ref{Table2} for all three regimes characterized by $S_{1}/L\approx
969: 0.3 $, $S_{1}\approx L$ and $S_{1}/L\approx 3$, respectively. 
970: \begin{table*}[tbp]
971: \centering%
972: \begin{minipage}{140mm}
973: \caption{Order of magnitude estimates for the inspiral rate $\dot{L}/L$,
974: angular precessional velocity $\Omega _{p}$ and tilt velocity $\dot{\protect%
975: \alpha}$ of the vectors $\mathbf{L}$ and $\mathbf{S}_{\mathbf{1}}$ with
976: respect to $\mathbf{J}$, represented for the three regimes with $L>S_{1}$, $%
977: L\approx S_{1}$ and $L<S_{1}$, characteristic in the domain of mass ratios $%
978: \protect\nu =0.3\div 0.03$. The numbers in brackets represent inverse time scales in seconds$^{-1}$, calculated for the typical
979: mass ratio $\protect\nu =10^{-1}$, post-Newtonian parameter $10^{-3}$, $%
980: 10^{-2}$ and $10^{-1}$, respectively and $m=10^{8}M_{\odot }$ (then $%
981: c^{3}/Gm=2\times 10^{-3}$ s$^{-1}$).}
982: \begin{center}
983: \begin{tabular}{crlrlrl}
984:    		& 
985: $L>S_{1}$ 	&
986: 		& 
987: $L\approx S_{1}$ & 
988: 		&
989: $L< S_{1}$  & \\
990: \hline\hline
991: $-\dot{L}/L $ 					 & 
992: $\quad\frac{32c^{3}}{5Gm}\varepsilon^{4}\eta $  &
993: $\left(\approx 10^{-15}\right) $ 		 & 
994: $\quad\frac{32c^{3}}{5Gm}\varepsilon ^{4}\eta $ &
995: $\left(\approx 10^{-11}\right) $ 		 & 
996: $\quad\frac{32c^{3}}{5Gm}\varepsilon ^{4}\eta $ &
997: $\left(\approx 10^{-7}\right) $ \\
998: $\quad\Omega _{p}$ 				 & 
999: $\quad\frac{2c^{3}}{Gm}\varepsilon ^{5/2}\eta $ &
1000: $\left(\approx 10^{-11}\right) $ 		 & 
1001: $\quad\frac{2c^{3}}{Gm}\varepsilon ^{5/2}\eta \frac{J}{L}$ &
1002: $\left(\approx 10^{-8}\frac{J}{L}\right) $ & 
1003: $\quad\frac{2c^{3}}{Gm}\varepsilon ^{3}$		 &
1004: $\left(\approx 10^{-5}\right) $ \\
1005: $\quad\frac{\dot{\alpha}}{\sin(\alpha+\beta)} $ 				 & 
1006: $\quad\frac{32c^{3}}{5Gm}\varepsilon ^{9/2}\frac{\eta}{\nu}$	 &
1007: $\left(\approx 10^{-16}\right) $ 		 & 
1008: $\quad\frac{32c^{3}}{5Gm}\varepsilon ^{9/2}\frac{\eta}{\nu}\frac{L^{2}}{J^{2}}$ &
1009: $\left(\approx 10^{-11}\frac{L^{2}}{J^{2}}\right)$	     & 
1010: $\quad\frac{32c^{3}}{5Gm}\varepsilon ^{7/2}\eta\nu$	     &
1011: $\left(\approx 10^{-8}\right) $
1012: \end{tabular}
1013: \end{center}
1014: \label{Table2}
1015: \end{minipage}
1016: \end{table*}
1017: 
1018: The numbers from the second line of Table \ref{Table2} demonstrate that for
1019: the chosen typical example the precession timescale can get as short as a
1020: day, going from 3000 years to three years to a day in the three columns
1021: above. This last stage is obviously quite close to the plunge. From the
1022: first line we can infer upper limits of how close the merger actually is, so
1023: 30 million years in Column 1, 300 years in Column 2, and a few months in
1024: Column 3. As the inspiral rate increases in time, rather than being a
1025: constant, these numbers are higher than the real values. The accuracy of the
1026: third estimate is further obstructed by the fact that after ISO the plunge
1027: follows, but as this comprises only a few orbits, the PN prediction can be
1028: considered relevant as an order of magnitude estimate. By multiplying the
1029: numbers in the third line with the precession timescales $\Omega _{p}^{-1}$
1030: we actually get the relevant tilt angle, varying from 2 arcsec during one
1031: precession ($6\times 10^{-4}$ arcseconds per year) at large separations to 3
1032: arcmin per precession (per day) close to the ISO.
1033: 
1034: We see, that the rate of precession and the tilt velocity become comparable
1035: in the $S_{1}\approx L$ epoch (in which $\varepsilon ^{1/2}\nu ^{-1}\approx
1036: 1 $) for 
1037: \begin{equation}
1038: \left[ \frac{16}{5}\sin \left( \alpha +\beta \right) \right] ^{-1/3}\frac{J}{%
1039: L}\approx \left( \varepsilon ^{2}\nu ^{-1}\right) ^{1/3}\approx \varepsilon
1040: ^{1/2}\approx \nu ~,  \label{transitional}
1041: \end{equation}%
1042: that is, for the chosen numerical example $\nu =10^{-1}$ and for the square
1043: bracket of order unity, this gives $J/L\approx 10^{-1}$ and the rate of $%
1044: \dot{\alpha}\approx \Omega _{p}\approx 10^{-9}$ (this is still $100$ times
1045: faster than the rate of inspiral).
1046: 
1047: The total angular momentum $J$ can be that small only if $\mathbf{L}$ and $%
1048: \mathbf{S_{1}}$ are almost perfectly anti-aligned, thus $\alpha +\beta
1049: \approx \pi -\delta $, $\delta \ll 1$, and~$L\approx S_{1}$. What does the
1050: condition for transitional precession (\ref{transitional}) mean in terms of
1051: their angle, how close should that be to the perfect anti-alignment? To
1052: answer the question we note that%
1053: \begin{equation}
1054: \frac{J}{L}=\frac{L\cos \alpha +S_{1}\cos \beta }{L}\approx \allowbreak
1055: \delta \sin \alpha ~,
1056: \end{equation}%
1057: Then, comparing with the estimates (\ref{albe}) and (\ref{transitional}) we
1058: conclude, that transitional precession can occur only if the deviation from
1059: the perfect anti-alignment is of the order of $\nu ^{3/2}$. This is a highly
1060: untypical case in galactic black hole binaries.
1061: 
1062: \subsection{The limits of validity}
1063: 
1064: One might seriously question whether pushing the values of the parameters
1065: beyond the range for which we use them would actually demonstrate that our
1066: approximations cannot possibly be correct. We would like to address the
1067: following two concerns specifically.
1068: 
1069: a) Is the high orbital angular momentum limit $L>>S_1$, obtained by a
1070: sufficient increase of the separation of the two black holes a correct limit?
1071: 
1072: b) Is the extreme mass limit $\nu \rightarrow 0$, for which the treatment of
1073: Section 5 may seem increasingly accurate, correct?
1074: 
1075: Concerning the limit (a): in any expansion with a small parameter one
1076: condition always holds: the parameter must be small, and in any PN expansion
1077: we can always reach a stage, for which the physics basis fails as
1078: inadequate, as other additional physical processes become dominant, or for
1079: which there is no observational support, despite the mathematical validity
1080: of the expansion.
1081: 
1082: From among the effects affecting the directions of the spins the dynamics
1083: discussed in Section 5 takes into account (1) the leading order conservative
1084: effect, given by the precession due to spin-orbit coupling and (2) the
1085: leading order dissipative effect due to gravitational radiation. Other
1086: conservative and dissipative effects are neglected, being weaker. Meaningful
1087: results can be traced from this model only when these assumptions hold. This
1088: implies that the post-Newtonian parameter $\varepsilon $ varies between $%
1089: 10^{-3}$ and $10^{-1}$, corresponding to orbital separations of 500
1090: Schwarzschild radii $=0.005$ pc to 5 Schwarzschild radii, given a $10^{8}$ $%
1091: M_{\odot }$ black hole. This is emphasized in the paragraph following Eq. (%
1092: \ref{rep}).
1093: 
1094: The "initial" and "final" phases in the dynamics described above therefore
1095: refer to a well-defined range of orbital separation, they are not arbitrary.
1096: The choice for the initial state is further justified by the discussion of
1097: Section 4. One cannot apply the dynamics discussed above at arbitrarily
1098: large distances, where the orbital momentum indeed would dominate, simply
1099: because the dynamics is no longer valid there. At larger separations the
1100: leading order dissipative effect is due to dynamical friction, thus the
1101: discussion of the previous two subsections does not apply.
1102: 
1103: Concerning (b): according to the summary presented in Table 1, there are
1104: three possibilities in the PN regime, where the dynamics discussed above
1105: holds: (i) from mass ratios ${\nu }$ from $1$ to $1/3$ the orbital angular
1106: momentum dominates throughout the whole range of separations between $0.005$
1107: pc and 5 Schwarzschild radii, as noted; (ii) from mass ratios $\nu $ from $%
1108: 1/3$ to $1/30$ initially the orbital angular momentum dominates over the
1109: spin, but their ratio is reversed at the final separation; (iii) from mass
1110: ratios $\nu $ smaller than $1/30$ the spin is dominant throughout the
1111: process.
1112: 
1113: Our claim is that the spin-flip is produced only if the total angular
1114: momentum (whose direction stays unchanged) initially is dominated by the
1115: orbital angular momentum, finally by the spin, thus only for the case (ii).
1116: 
1117: In the dynamics presented in the previous two subsections we have neglected
1118: the second spin. As even for the highest mass ratio $\nu =1/3$ in the regime
1119: (ii) the second spin is 1 order of magnitude smaller than the leading spin,
1120: we consider this assumption justified. With decreasing mass ratios it
1121: becomes increasingly accurate to neglect the second spin, as according to
1122: Eq. (7) the ratio of the spins goes with $\nu ^{2}$.
1123: 
1124: However, not all results of the previous subsection become increasingly
1125: accurate with a decreasing mass ratio. We emphasize, what is different in
1126: case (iii) as compared to (ii). The difference is in the initial conditions,
1127: which allow to obtain a spin-flip in case (ii), but not in (iii).
1128: Mathematically, the difference between these two cases can be seen from Eq.
1129: (33), showing that the angular tilt velocity of the dominant spin scales
1130: with $\nu ^{2}$. For the extreme mass ratios $\nu <1/30$, e.g., Table 1, the
1131: spin dominates over the orbital angular momentum throughout the whole PN
1132: regime. Therefore, the ratio $S_{1}/J$ is of order unity. With decreasing $%
1133: \nu $ however the change in the direction of the spin, represented by $\dot{%
1134: \alpha}$ goes fast to zero, thus no spin-flip is produced in the PN regime
1135: for extreme mass ratios.
1136: 
1137: At the end of this subsection we derive an analytical expression relating $%
1138: \alpha $ to the conserved $\alpha +\beta $, the evolving PN parameter $%
1139: \varepsilon $ and the mass ratio $\nu $. From the Figure \ref{Fig1} $J=L\cos
1140: \alpha +S_{1}\cos \beta $. By introducing the angle $\alpha +\beta $ and
1141: employing the estimate (\ref{S1L}) we obtain 
1142: \begin{equation}
1143: \frac{J}{L}\approx \left[ 1+\varepsilon ^{1/2}\nu ^{-1}\cos \left( \alpha
1144: +\beta \right) \right] \cos \alpha +\varepsilon ^{1/2}\nu ^{-1}\sin \left(
1145: \alpha +\beta \right) \sin \alpha ~.
1146: \end{equation}%
1147: Inserting this into the second expression (\ref{albe}) and rearranging we
1148: find%
1149: \begin{equation}
1150: \frac{\sin 2\alpha }{1+\cos 2\alpha }\approx \frac{\sin \left( \alpha +\beta
1151: \right) }{\varepsilon ^{-1/2}\nu +\cos \left( \alpha +\beta \right) }~.
1152: \end{equation}%
1153: For an initial configuration of $0.005$ pc (such that $\varepsilon \equiv
1154: \varepsilon ^{\ast }=10^{-3}$) and mass ratio $\nu =10^{-1}$, the initial
1155: misalignment between $\mathbf{L}$ and $\mathbf{J}$ is $\alpha
1156: _{initial}\approx 18^{\circ },~10^{\circ },~0^{\circ }$ for the dominant
1157: spin in the plane of orbit, spanning $45^{\circ }$ with the plane of orbit
1158: and perpendicular to the plane of orbit (such that $\alpha +\beta =90^{\circ
1159: },~45^{\circ },~0^{\circ }$), respectively. Then $\beta _{initial}=72^{\circ
1160: },~35^{\circ },~0^{\circ }$. For the same mass ratio and relative
1161: configurations, the angle $\alpha $ at the end of the PN epoch (at $%
1162: \varepsilon =10^{-1}$) becomes $\alpha _{final}\approx 73^{\circ
1163: },~35^{\circ },~0^{\circ }$,\ respectively. This can be translated into a
1164: misalignment between $\mathbf{S}_{\mathbf{1}}$ and $\mathbf{J}$ of $\beta
1165: _{final}=17^{\circ },~10^{\circ },~0^{\circ }$, and a spin-flip of $\Delta
1166: \beta =55^{\circ },~25^{\circ },~0^{\circ }$, respectively.
1167: 
1168: \subsection{Summary}
1169: 
1170: In the typical range of mass ratios $\nu =0.03\div 0.3$ the initial
1171: condition $L>S_{1}$ is always transformed into $S_{1}<L$, but the transition
1172: is very rarely accompanied by the so-called transitional precession. In all
1173: other cases the precession is simple. As the precession angle of the
1174: dominant spin is decreasing in time from the given initial value to a small
1175: value, the precessional cone becomes narrower in time. At the end of the
1176: inspiral phase the dominant spin $\mathbf{S}_{\mathbf{1}}$ will point
1177: roughly along $\mathbf{J.}$ This means that a spin-flip has occurred during
1178: the post-Newtonian evolution, already in the inspiral phase of the merger.
1179: On the other hand, as the inspiral phase ends with $L<S_{1}$, irrespective
1180: of what happens in the next phases, during the plunge and ring-down, $L$ is
1181: not high enough to cause additional significant spin-flip.
1182: 
1183: For smaller mass ratios (for extreme mass ratio mergers) the larger spin
1184: already dominates the total angular momentum from the beginning of the
1185: inspiral, thus no spin-flip will occur by the mechanism presented here.
1186: Alternatively, from the second expression (\ref{alphadotepeta}) one can see
1187: that the rate of tilt of the spin decreases with $\nu ^{2}$, thus it goes
1188: fast to zero in the extreme mass ratio case. However, as we argued in
1189: Section 2, such mass ratios are less typical for galactic central SMBH
1190: mergers. This also shows that an infalling particle will not change the spin
1191: of the supermassive black hole.
1192: 
1193: For the (again untypical) equal mass SMBH mergers the orbital angular
1194: momentum stays dominant until the end of the inspiral phase. In this case,
1195: however the possibility remains open to have a spin-flip later on, during
1196: the plunge.
1197: 
1198: \section{Discussion}
1199: 
1200: The considerations from this paper lead to the following time sequence for
1201: the transient feeding of a SMBH including a merger with another SMBH.
1202: 
1203: \textit{First:} Two galaxies with central black holes approach each other to
1204: within a distance where dynamical friction keeps them bound, spiraling into
1205: each other. If there is cool gas in either one, it can begin to form stars
1206: rapidly, along tidal arms. The galactic central supermassive SMBH binary
1207: influences gas dynamics and star formation activity also in the nuclear gas
1208: disk, due to various resonances between gas motion and SMBH binary motion
1209: (Matsui et al. 2006), creating some characteristic structures, such as
1210: filament structures, formation of gaseous spiral arms, and small gas disks
1211: around SMBHs. If either galaxy happens to have radio jets, then due to the
1212: orbital motion, these jets get distorted and form the Z-shape (Gopal-Krishna
1213: et al. 2003, Zier 2005).
1214: 
1215: \textit{Second: }The central regions in each galaxy begin to act as one
1216: unit, in a sea of stars and dark matter of the other galaxy. During this
1217: phase, as the cool gas from the other partner typically has low angular
1218: momentum with respect to the receiving galaxy, the central region can be
1219: stirred up, and produce a nuclear starburst (Toomre and Toomre 1972). The
1220: central black hole can get started to be fed at a high rate, but its
1221: emission will be submerged in all the far-infrared emission from the gas and
1222: dust heated by the massive stars produced in the starburst. In this case,
1223: there is dynamical friction, which can act so as to select certain
1224: symmetries, such as corotation, counter rotation, or rotation at $90^{\circ
1225: } $ (as in NGC2685, a polar ring galaxy; Richter et al. 1994).
1226: 
1227: \textit{Third: }The black holes begin to lose orbital angular momentum due
1228: to the interaction with the nearby stars (Zier and Biermann 2001, 2002). \
1229: Other mechanisms for angular momentum loss are also known (Sesana et al.
1230: 2006, 2007a,b, Alexander 2007, Hayasaki 2008). The two black holes approach
1231: each other to that critical distance where the interaction with the stars
1232: and the gravitational radiation remove equivalent fractions of the orbital
1233: angular momentum. Then, as shown in this paper, the spin axes tumble and
1234: precess. This phase can be identified with the apparent superdisk, as the
1235: rapidly precessing jet produces the hydrodynamic equivalent of a powerful
1236: wind, by entraining the ambient hot gas, pushing the two radio lobes apart
1237: and so giving rise the a broad separation (Gopal-Krishna et al. 2003, 2007,
1238: Gopal-Krishna \& Wiita 2006). Gopal-Krishna \& Wiita (2006) emphasize the
1239: apparent asymmetry, which we propose to attribute to line-of-sight effects
1240: and the distortion due to the recent merger. The base of the radio structure
1241: is so broad and so asymmetric, that the central AGN will appear to be offset
1242: from the projected center of gap. The recent arguments of Worrall et al.
1243: (2007) seem to be consistent with this point of view. The spin direction of
1244: the combined two black holes is preserved, although the strength of the spin
1245: decreases. As during simple precession the total angular momentum shrinks
1246: considerably, but its direction is conserved, on the other side the
1247: magnitude of the spin stays constant, this means that the orbital angular
1248: momentum shrinks. For comparable mass binaries it will be still higher than
1249: the spin at ISO (therefore the dynamics below ISO, which can be analyzed
1250: only numerically, should be responsible for any spin-flip in the comparable
1251: masses case). For extreme mass ratio binaries the result of the shrinkage of
1252: the orbital angular momentum is $L<S$ at ISO. Therefore, the spin at ISO
1253: should be roughly aligned with the direction of $\mathbf{J}=\mathbf{L}+%
1254: \mathbf{S}$, which (as initially $\mathbf{L}$ was dominant), is close to the
1255: direction of the initial $\mathbf{L}$.
1256: 
1257: In certain cases, especially for equal masses of the two black holes, a
1258: strong recoil has been found (Gonz{\'{a}{}}lez et al. 2007a, b). However, as
1259: we noted earlier, the equal mass ratio is untypical.
1260: 
1261: \textit{Fourth:} The two black holes actually merge, and the merged black
1262: hole keeps the spin axis from the orbital angular momentum of the previously
1263: existing binary, whenever the mass ratio is relatively large. In the case
1264: that the mass ratio is between $1:1$ and $1:3$, then even at the innermost
1265: stable orbit a substantial fraction of the orbital angular momentum can
1266: survive, possibly leading to a spin-flip later on. This very short phase
1267: should be accompanied by extreme emission of low-frequency gravitational
1268: waves. The final stage in this merger leads to a rapid increase in the
1269: frequency of the waves, called \textquotedblleft chirping", but this
1270: chirping will depend on the angles involved. The angle between the orbital
1271: spin of the combined two black holes, and intrinsic spin of the more massive
1272: black hole influences the highest frequency of the chirp; for a large angle
1273: this frequency will be lower than for a small angle between the two spins.
1274: Whether there is another observable feature, such as the induced decay of
1275: heavy dark matter particles, from the merger of the two black holes at that
1276: event such as speculated by Biermann \& Frampton (2006) is not clear at this
1277: time.
1278: 
1279: \textit{Fifth:} Now the newly oriented more massive merged black hole starts
1280: its accretion disk and jet anew, boring a new hole for the jets through its
1281: environment. This stage can be identified perhaps with giga hertz peaked
1282: radio sources (GPS). If the new jet points at the observer, then 3C147 may
1283: be one example (Junor et al. 1999).
1284: 
1285: \textit{Sixth:} The newly oriented jets begin to show up over some kpc, and
1286: this corresponds to the X-shaped radio galaxies, while the old jets are
1287: fading but still visible. This also explains many of the compact steep
1288: spectrum sources, with disjoint directions for the inner and outer jets.
1289: 
1290: \textit{Seventh: }The old jets have faded, and are at most visible in the
1291: low radio frequency bubbly structures, such as seen for the Virgo cluster
1292: region around M87 (Owen et al. 2000). The feeding is slowing down, and there
1293: is no longer an observable accretion disk, but probably only an
1294: advection-dominated disk. However, a powerful jet is still there, although
1295: below or even far below the maximal power. The feeding is still from the
1296: residual material stemming from the merger.
1297: 
1298: \textit{Eighth:} The feeding of the black hole is down to catching some gas
1299: out of a common red giant star wind as presumably is happening in our
1300: Galactic center. This stage seems to exist for all black holes, even at very
1301: low levels of activity (e.g., Perez-Fournon \& Biermann 1984, Elvis et al.
1302: 1984, Nagar et al. 2000).
1303: 
1304: If this concept described here is true, then the superdisk radio galaxies
1305: should have large outer distortions in their radio images, that may be
1306: detectable at very high sensitivity, as they should correspond to recently
1307: active Z-shaped sources. Also, the superdisk should be visible in X-rays,
1308: although if the cooling is efficient the temperature may be relatively low.
1309: Table 2 suggests that the merger is imminent, if the precession of the jet
1310: is measurable within a few years, and the opening angle of the precession is
1311: much narrower than the wind cone, reflecting the earlier longer time
1312: precession (see Gopal-Krishna et al. 2007). Therefore, with very sensitive
1313: radio interferometry it might be possible to detect the underlying jet
1314: despite its rapid precession, although immediately before the actual merger
1315: the feeding of the jet will be turned off.
1316: 
1317: As more and more pieces of evidence suggest that AGNs are the sources of
1318: ultra-high energy cosmic rays (Biermann \& Strittmatter 1987, Biermann et
1319: al. 2007) we need to ask what we could learn next. Clearly, after a
1320: spin-flip, the new relativistic jet bores through a new environment, with
1321: lots of gas, and so suffers a strong decelerating shock. In such a shock
1322: particles are accelerated to maximal energies, and at the same time, as they
1323: leave the shock region interact with all that interstellar gas. Therefore,
1324: such sites are primary sources for any new particles, such as high energy
1325: neutrinos (Becker et al. 2007). Such discoveries may well be possible long
1326: before we detect the low frequency gravitational waves from the black hole
1327: merger. As at such high energy neutrinos travel straight across the
1328: universe, and suffer little loss other than from the adiabatic expansion of
1329: the universe, the black holes resulting from a merger of two black holes,
1330: with subsequent spin-flip, will be primary targets for searches for
1331: ultra-high energy neutrinos, and perhaps other photos and particles at
1332: extreme energies.
1333: 
1334: \section{Concluding Remarks}
1335: 
1336: Whereas it has been questioned in the past whether the central SMBHs of
1337: merging galaxies will be able to actually merge or their approach will stall
1338: (due to the process of loss-cone depletion) at a distance where dissipation
1339: through gravitational radiation is not yet efficient (for a review of these
1340: considerations see Merritt \& Miloslavljevi\'{c} 2005), the role of the
1341: dynamical friction as bringing close the SMBHs to the transition radius,
1342: from where gravitational radiation undertakes the control of the dissipative
1343: process has been recently confirmed (Zier 2006) and also complementary
1344: mechanisms were proposed (Alexander 2007, Sesana et al. 2006, 2007a,b,
1345: Hayasaki 2008). The space mission LISA is predicted to detect the merger of
1346: SMBHs. The statistical arguments of Rottmann (2001), using radio
1347: observations, suggest that all strong central activity in galaxies may
1348: involve a merger of two black holes. Therefore, we have assumed in this
1349: paper that whenever galaxies merge, so will do their central SMBHs. Even if
1350: there would be exceptions under this rule, this would reflect only in the
1351: inclusion of an overall factor $\lesssim 1$ in the number of mergers of
1352: SMBHs as compared to the number of mergers of galaxies, derived in Section
1353: 2, which would not affect the mass ratio estimates of our paper.
1354: 
1355: Guided by reasonable and simple assumptions we have shown that binary
1356: systems of \textit{SMBH binaries formed by galaxy mergers typically have a
1357: mass ratio range between }$1/3$\textit{\ and }$1/30$. Following this, we
1358: have proven that for the typical mass ranges a combination of the SO
1359: precession and gravitational radiation driven dissipation produces \textit{%
1360: the spin-flip of the dominant black hole already in the inspiral phase},
1361: except for the particular configuration of the spin perpendicular to the
1362: orbital plane. During this process the magnitude of the spin is unchanged,
1363: therefore \textit{the merger of a high spin} (and high rotation parameter) 
1364: \textit{black hole with the smaller black hole results in a similar high
1365: spin state} at the end of the inspiral phase. These are the main results of
1366: our paper.
1367: 
1368: There is a related discussion undergoing in the literature, whether the high
1369: spin of SMBHs is produced by prolonged accretion phases or by frequent
1370: mergers. Even a scenario, where the SMBHs have typically low spin (King \&
1371: Pringle 2006) was advanced, based on the assumption of short periods of
1372: small accretion from random directions. Hughes \& Blandford (2003),
1373: extrapolating results from $\nu =q^{-1}\ll 1$ binaries to comparable masses,
1374: have shown that mergers spin-down black holes. Volonteri et al. (2005) have
1375: studied the distribution of SMBH spins under the combined action of
1376: accretion and mergers, and found that the dominant spin-up effect is by gas
1377: accretion. Recently, Berti \& Volonteri (2008) have considered the problem
1378: of mergers by taking into account improvements in the numerical general
1379: relativistic methods (Pretorius 2007), and a recent semianalytical formula,
1380: which gives the final spin in terms of the initial dimensionless spins, mass
1381: ratio, and relative angles of orbital angular momentum and spins (Rezzolla
1382: et al. 2008a,b,c, Barausse \& Rezzolla 2009). They have found that mergers
1383: can result in a high spin end state only if the dominant spin is aligned
1384: with the orbital angular momentum of the system (thus the smaller mass
1385: orbits in the equatorial plane of the larger). Their considerations extend
1386: from comparable masses to mass ratios of $1/10$. However, Berti and
1387: Volonteri (2008) neglected the angular momentum exchange and transport
1388: between black hole, jet, and inner accretion disk by magnetic fields (see,
1389: e.g., Blandford 1976); this may modify or even sharpen the conclusions.
1390: 
1391: We can add three remarks to this discussion. First, we have shown by
1392: analytical means, that for the typical mass ratio range the inspiral phase
1393: ends with a considerably lower value of the orbital angular momentum
1394: compared to the spin (see the last picture in Figure \ref{Fig1}). A
1395: heuristic argument then shows that such a small angular momentum could not
1396: significantly change the \textit{direction} of the spin during the next
1397: phases of the merger. Apart from this small orbital angular momentum, the
1398: problem being axially symmetric, \textit{we do not expect significant
1399: further spin-flip due to gravitational radiation in the last stages of the
1400: inspiral}.
1401: 
1402: Second, the configuration of \textit{orbital angular momentum aligned with
1403: the dominant spin is not a preferred one} in the gravitational radiation
1404: dominated post-Newtonian regime. It is not clear yet whether such an
1405: alignment could be the by-product of previous phases of the inspiral, when
1406: dynamical friction (Zier \& Biermann 2001), three-body interactions (Sesana
1407: et al. 2006, 2007a,b), relaxation processes due to cloud-star interactions
1408: (Alexander 2007), three disk model accretion (Hayasaki 2008), and other
1409: possible mechanisms occur. Since the stellar system is often slightly
1410: flattened, differential dynamical friction could produce the near alignment
1411: necessary to allow very high spin after a merger.
1412: 
1413: Third, the \textit{magnitude} of the spin is practically unchanged in the
1414: inspiral phase, discussed here. This is because the loss in the spin vector
1415: by gravitational radiation, a second PN order effect, calculated from the
1416: Burke-Thorne potential (Burke 1971), is perpendicular to the spins, yielding
1417: another precessional effect (Gergely et al. 1998c). Below ISO this estimate
1418: should break down, as indicated by numerical simulations reporting on
1419: various fractions of the spin radiated away. In this context we want to
1420: emphasize the \textit{unchanged magnitude of the spin during the inspiral},
1421: as important initial data for the numerical evolution during the plunge and
1422: ring-down.
1423: 
1424: We also mention here the results of the numerical relativity community
1425: showing a considerable recoil of the merged SMBH in particular cases, mostly
1426: for equal masses and peculiar configurations of the angular momenta (Br{\"{u}%
1427: }gmann 2008, Gonzalez et al. 2007a, b, Koppitz et al. 2007). It has also
1428: been shown that the recoil regulates the SMBH mass growth, as the SMBH
1429: wanders through the host galaxy for $10^{6}\div 10^{8}$ years (Blecha and
1430: Loeb 2008). According to the empirical formula of Campanelli et al. (2007a,
1431: see also Lousto \& Zlochower 2009) the recoil velocity scales with $%
1432: q^{-2}/\left( 1+q^{-1}\right) ^{2}\left( 1+q^{-1}\right) $, which for $%
1433: q^{-1}=\nu \ll 1$ reduces to a scaling with $q^{-2}$. Therefore, \textit{we
1434: do not expect significant recoil effects in the typical mass ratio range of
1435: the SMBH mergers}.
1436: 
1437: We suggest that the precessional phase of the merger of two black holes,
1438: occurring prior to the spin-flip, is visible as a superdisk in radio
1439: galaxies (Gopal-Krishna et al. 2007). The \textit{precessing jet appears as
1440: a superwind separating the two radio lobes} in the final stages of the
1441: merger. According to our model such radio galaxies are candidates for
1442: subsequent SMBH mergers. Further observations and theoretical work may be
1443: capable of identifying such candidates likely to merge, and determine the
1444: timescale for this to happen. The restart of powering a relativistic jet
1445: (after the spin-flip and merging) will produce ultra-high energy hadrons,
1446: neutrinos and other particles.
1447: 
1448: Based on the estimates given in Table 2 for the precessional and inspiral
1449: timescales, we can say the following. If we were to observe a precession
1450: timescale of three years in a superdisk radio galaxy, we could confidently
1451: predict a plunge in about 300 years, which should be observable. Faster
1452: precession timescales would take some effort to identify. However, if we
1453: were able to even identify a precession timescale of days to weeks, then the
1454: plunge would be predicted to happen a few months to a few years thence:
1455: powerful gravitational waves at very low frequency would then be emitted.
1456: 
1457: The picture developed here differs from that in Wilson \& Colbert (1995) in
1458: that we do not identify just the rare mergers of two massive black holes of
1459: about equal masses with radio galaxies and radio quasars. We intend to
1460: revisit the interactions with the stars (Zier et al. 2009, in preparation),
1461: discuss the spin of the black holes in another work (Kov\'{a}cs et al. 2009,
1462: in preparation) developed from Du\c{t}an \& Biermann (2005), finally to work
1463: out quantitatively the relation of the merger of black holes and the
1464: statistics of radio galaxies (Gopal-Krishna et al. 2009, in preparation).
1465: 
1466: \section{Acknowledgements}
1467: 
1468: We are grateful for discussions with Gopal-Krishna and C. Zier. P.L.B.
1469: acknowledge further discussions with J. Barnes, B. Br{\"{u}}gmann, and G. Sch%
1470: {\"{a}}fer. L.\'{A}.G. was successively supported by OTKA grants 46939,
1471: 69036, the J\'{a}nos Bolyai Grant of the Hungarian Academy of Sciences, the
1472: London South Bank University Research Opportunities Fund and the Pol\'{a}nyi
1473: Program of the Hungarian National Office for Research and Technology (NKTH).
1474: Support for P.L.B. was from the AUGER membership and theory grant 05 CU 5PD
1475: 1/2 via DESY/BMBF and VIHKOS. The collaboration between the University of
1476: Szeged and the University of Bonn was via an EU Sokrates/Erasmus contract.
1477: 
1478: \begin{thebibliography}{999}
1479: \bibitem{1240} Alexander, T., in \textit{2007 STScI Spring Symp.: Black
1480: Holes"}, eds, M. Livio \& A.M. Koekemoer, (Cambridge, Cambridge University
1481: Press), in press (arXiv:0708.0688)
1482: 
1483: \bibitem{1244} Antonucci, R.R.J., Miller, J.S., \textit{Astrophys. J.} 
1484: \textbf{297}, 621 - 632 (1985) 
1485: %  title:  Spectropolarimetry and the nature of NGC 1068
1486: % rediscovery of the torus
1487: 
1488: \bibitem{1253} Apostolatos T.A., \textit{Phys. Rev. D }\textbf{52}, 605
1489: (1995) 
1490: %  title: Search templates for gravitational waves from precessing, inspiraling binaries 
1491: 
1492: \bibitem{1256} Apostolatos T.A., \textit{Phys. Rev. D }\textbf{54}, 2438
1493: (1996) 
1494: % title: Influence of spin-spin coupling on inspiraling compact binaries with M1=M2 and S1=S2 
1495: 
1496: \bibitem{1249} Apostolatos T.A., Cutler C., Sussman G.J., Thorne K.S., 
1497: \textit{Phys. Rev. D} \textbf{49}, 6274 (1994) 
1498: % title: Spin-induced orbital precession and its modulation of the gravitational waveforms from merging binaries
1499: 
1500: \bibitem{} Barausse, E. Rezzolla, L., arXiv:0904.2577V1 [gr-qc] (2009) 
1501: % title: Predicting the direction of the final spin from the coalescence of two black holes
1502: 
1503: \bibitem{1259} Barker\ B.M., O'Connell R.F., \textit{Phys. Rev. D }\textbf{12%
1504: }, 329 (1975)
1505: 
1506: \bibitem{1262} Barker B.M., O'Connell R.F., \textit{Gen. Relativ. Gravit.} 
1507: \textbf{2}, 1428 (1979)
1508: 
1509: \bibitem{1269} %  Barnes, J.E. ... reviews
1510: Barnes, J.E., in \textit{Proc. of the 4th Sci. Meet. of the Span. Astron.
1511: Soc. (SEA)} 2000, Highlights of Span. Astrophys. II. ed. J. Zamorano, J.
1512: Gorgas, \& J. Gallego (Dordrecht: Kluwer), 85 
1513: %  title:  Dynamics of galaxy interactions (Invited Review)
1514: 
1515: \bibitem{1265} Barnes, J.E., Hernquist, L., \textit{Annual Rev. of Astron.
1516: \& Astrophys.} \textbf{30}, 705 (1992) 
1517: %  title:  Dynamics of interacting galaxies
1518: 
1519: \bibitem{1275} Barth A.J., Greene J.E., Ho L.C., \textit{Astrophys. J.
1520: Letters} \textbf{619}, L151 (2005) 
1521: % title: Dwarf Seyfert 1 Nuclei and the Low-Mass End of the MBH-$\sigma$
1522: % Relation ; astro-ph/0412575
1523: 
1524: \bibitem{1280} Becker J.K., Gro\ss\ A., M\"{u}nich K., Dreyer J., Rhode W.,
1525: Biermann P.L., \textit{Astropart. Phys.} \textbf{28}, 98 (2007) 
1526: %Astrophysical implications of high energy neutrino limits,
1527: %; astro-ph/0607427
1528: 
1529: \bibitem{1285} Benson A.J., D\v{z}anovi\'{c} D., Frenk C.S., Sharples R., 
1530: \textit{Mon. Not. Roy. Astron. Soc. }\textbf{379}, 841-866 (2007) 
1531: %  title:  "Luminosity and stellar mass functions of disks and spheroids in the SDSS and the supermassive black hole mass function", astro-ph/0612719v2
1532: 
1533: \bibitem{1289} %  BBH-h-63076.web
1534: Berczik P., Meritt D., Spurzem R., \textit{Astrophys. J. Letters} \textbf{633%
1535: }, 680 - 687 (2005) 
1536: %  title:  Long term evolution of massive black hole binaries II. Binary 
1537: %  evolution in low density galaxies
1538: %  testing loss cone theory... when scaled to real galaxies, our results suggest %  that massive black hole binaries in gas poor nuclei would be unlikely to 
1539: %  reach gravitational wave coalescence in a Hubble time
1540: 
1541: \bibitem{1297} %  BBH-g-20398.web
1542: Berczik P., et al., \textit{Astrophys. J. Letters} \textbf{642}, L21 - L24
1543: (2006)
1544: 
1545: \bibitem{1304} Berti E., Volonteri M., \textit{Astrophys. J.} \textbf{684},
1546: 822 (2008) %arXiv:0802.0025v2 [astro-ph] (2008)
1547: %  title:  Cosmological black hole spin evolution by mergers and accretion, 
1548: 
1549: \bibitem{1307} Biermann P.L., Strittmatter P.A., \textit{Astrophys. J.} 
1550: \textbf{322}, 643 (1987) 
1551: % title: Synchrotron emission from shockwaves in active galactic nuclei
1552: 
1553: \bibitem{1311} Biermann P.L., Strom R.G., Falcke H., \textit{Astron. \&
1554: Astroph.} \textbf{302}, 429 (1995) 
1555: % title: Cosmic rays V. The nonthermal radioemission of the old nova GK
1556: % Per - a signature of hadronic interactions?
1557: %; astro-ph/9508102
1558: 
1559: \bibitem{1317} Biermann P.L., Chirvasa M., Falcke H., Markoff S., Zier Ch.,
1560: invited review at the Paris Conference on Cosmology, June 2000, in
1561: Proceedings, Eds. N. Sanchez, H. de Vega, p. 148 - 164 (2005);
1562: astro-ph/0211503 
1563: % title: Single and binary black holes and their active environment
1564: 
1565: \bibitem{1323} Biermann P.L., Frampton P.H., \textit{Physics Letters B} 
1566: \textbf{634}, 125 - 129 (2006) 
1567: % title: Ultra high energy cosmic rays from sequestered X bursts
1568: %; astro-ph/0512188
1569: 
1570: \bibitem{} Biermann P.L., Hall D.S., \textit{Astron. \& Astroph.} \textbf{27}%
1571: , 249 - 253 (1973). 
1572: %A model to explain alternate period changes in Algol-like binaries, 
1573: 
1574: \bibitem{1328} Biermann P. L., Isar P.G., Mari\c{s} I.C., Munyaneza F., Ta%
1575: \c{s}c\u{a}u O., "Origin and physics of the highest energy cosmic rays: What
1576: can we learn from Radio Astronomy ?\textit{"}, invited lecture at the Erice
1577: meeting June 2006, editors M.M. Shapiro, T. Stanev, J.P. Wefel, World
1578: Scientific, p. 111 (2007); astro-ph/0702161
1579: 
1580: \bibitem{} Biermann P. L., Becker J. K., Caramete A., Curu\c{t}iu L., Engel
1581: R., Falcke H., Gergely L. \'{A}., Isar P. G., Mari\c{s} I. C., Meli A.,
1582: Kampert K.-H., Stanev T., Ta\c{s}c\u{a}u O., Zier C., "Active Galactic
1583: Nuclei: Sources for ultra high energy cosmic rays?", invited review for the
1584: Proceedings of the CRIS 2008 - Cosmic Ray International Seminar: Origin,
1585: Mass, Composition and Acceleration Mechanisms of UHECRs, Malfa, Italy, Ed. A
1586: Insolia, Elsevier 2009; arXiv: 0811.1848v3 [astro-ph]
1587: 
1588: \bibitem{1334} Binney J., Tremaine S., \textit{Galactic Dynamics}, Princeton
1589: University Press (1987)
1590: 
1591: \bibitem{1337} Blanchet L., Buonanno A., Faye G., \textit{Phys. Rev.} 
1592: \textbf{D} \textbf{74}, 104034 (2006); Erratum-ibid. \textbf{75}, 049903
1593: (2007)
1594: 
1595: \bibitem{} Blandford R.D. \textit{Month. Not. Roy. Astr. Soc.} \textbf{176},
1596: 465 (1976)
1597: 
1598: \bibitem{1340} Blecha L., Loeb A., \textit{Month. Not. Royal Astron. Soc.} 
1599: \textbf{390}, 1311 (2008) %  arXiv: 0805.1420 [astro-ph]
1600: 
1601: \bibitem{1343} Bouwens R.J., Illingworth G.D., \textit{Nature} \textbf{443},
1602: 189 - 192 (2006) 
1603: %  title: Rapid evolution of the most luminous galaxies during the first 
1604: %  900 million years
1605: %  The first 900 million years (Myr) to redshift z ~ 6 (the first seven per cent %  of the age of the Universe) remains largely unexplored for the formation of 
1606: %  galaxies. Large samples of galaxies have been found at z ~ 6 (refs 1-4) but 
1607: %  detections at earlier times are uncertain and unreliable. It is not at all 
1608: %  clear how galaxies built up from the first stars when the Universe was about %  300Myr old (z ~ 12-15) to z ~ 6, just 600Myr later. Here we report the 
1609: %  results of a search for galaxies at z ~ 7-8, about 700Myr after the Big Bang, %  using the deepest near-infrared and optical images ever taken. Under 
1610: %  conservative selection criteria we find only one candidate galaxy at z ~ 7-8, %  where ten would be expected if there were no evolution in the galaxy 
1611: %  population between z ~ 7-8 and z ~ 6. Using less conservative criteria, there %  are four candidates, where 17 would be expected with no evolution. This 
1612: %  demonstrates that very luminous galaxies are quite rare 700Myr after the Big %  Bang. The simplest explanation is that the Universe is just too young to have %  built up many luminous galaxies at z ~ 7-8 by the hierarchical merging of 
1613: %  small galaxies.
1614: 
1615: \bibitem{1357} %  Br{\"u}gmann, B., .. 
1616: Br{\"{u}}gmann B., Gonzalez J., Hannam M., Husa S., Sperhake U., \textit{%
1617: Phys. Rev. }D \textbf{77}, 124047 (2008) 
1618: %  title:  Exploring black hole superkicks
1619: %  Recent calculations of the recoil velocity in black-hole binary mergers have %  found kick velocities of $\approx2500 $km/s for equal-mass binaries with 
1620: %  anti-aligned initial spins in the orbital plane. In general the dynamics of 
1621: %  spinning black holes can be extremely complicated and are difficult to 
1622: %  analyze and understand. In contrast, the ``superkick'' configuration is an 
1623: %  example with a high degree of symmetry that also exhibits exciting physics. 
1624: %; arXiv:0707.0135 
1625: 
1626: \bibitem{1367} Brunthaler A., Reid M.J., Falcke H., Greenhill L.J., Henkel
1627: C., \textit{Science} \textbf{307}, 1440 - 1443 (2005) 
1628: %  title:  The Geometric Distance and Proper Motion of the Triangulum Galaxy 
1629: %  (M33)
1630: %; astro-ph/0503058 
1631: 
1632: \bibitem{1373} Burke W.L., \textit{J. Math. Phys.} \textbf{12}, 401 (1971)
1633: 
1634: \bibitem{1375} %  BBH-d-21515.web
1635: Campanelli M., Lousto C.O., Zlochower Y., Merritt D. \textit{Astrophys. J.} 
1636: \textbf{659}, L5 (2007a) 
1637: % title: Merger Recoils and Spin Flips From Generic Black-Hole Binaries
1638: 
1639: \bibitem{1380} %  BBH-e-PRD_75_064030
1640: Campanelli M., Lousto C.O., Zlochower Y., Krishnan B., Merritt D., \textit{%
1641: Phys. Rev.} \textbf{D} \textbf{75}, 0640030 (2007b) 
1642: %  title:  Spin flips and precession in black hole binary mergers
1643: % numerical experiments; verifying the spin-flip concept
1644: 
1645: \bibitem{1386} Chini R., Kreysa E., Biermann P.L., \textit{Astron. \&
1646: Astroph.} \textbf{219}, 87-97 (1989a) 
1647: % title: The nature of radio-quiet quasars
1648: 
1649: \bibitem{1389} Chini R., Biermann P.L., Kreysa E., Gem{\"u}nd H.-P., \textit{%
1650: Astron. \& Astroph. Letters} \textbf{221}, L3 - L6 (1989b). 
1651: % title: 870 and 1300 micron observations of radio quasars
1652: 
1653: \bibitem{1393} Chirvasa M., Diploma thesis: "Gravitational Waves during the
1654: mergers of rotating black holes", Bonn Univ. (2001) 
1655: % title: Gravitational Waves during the mergers of rotating black holes
1656: 
1657: \bibitem{1397} Donea A.C., Biermann P.L., \textit{Astron. \& Astroph.} 
1658: \textbf{316}, 43 (1996) 
1659: % title: The symbiotic system in quasars: black hole, accretion disk and
1660: % jet
1661: %; astro-ph/9602092
1662: 
1663: \bibitem{1403} Du\c{t}an I., Biermann P.L., in the proceedings of the
1664: International School of Cosmic Ray Astrophysics (14th course): "Neutrinos
1665: and Explosive Events in the Universe", Ed. T. Stanev, published by Springer,
1666: Dordrecht, The Netherlands, p.175 (2005), astro-ph/0410194 
1667: %  title:  The Efficiency of Using Accretion Power of Kerr Black  Holes,
1668: 
1669: \bibitem{1409} Elvis M., Soltan A., Keel W.C., \textit{Astrophys. J.} 
1670: \textbf{283}, 479 - 485 (1984) 
1671: % title: Very low luminosity active galaxies and the X-ray background
1672: 
1673: \bibitem{1413} Faber S.M., Tremaine S., Ajhar E.A., et al. \textit{Astron. J.%
1674: } \textbf{114}, 1771 (1997) 
1675: % title: The Centers of Early-Type Galaxies with HST. IV. Central Parameter
1676: % Relations.
1677: % complete list of authors: Faber, S. M., Tremaine, S., Ajhar, E.A.,
1678: % Byun, Yong-Ik; Dressler, Alan; Gebhardt, Karl; Grillmair, Carl; Kormendy,
1679: % John; Lauer, Tod R.; Richstone, Douglas
1680: %; astro-ph/9610055
1681: 
1682: \bibitem{1422} Falcke H., Biermann P.L., \textit{Astron. \& Astroph.} 
1683: \textbf{293}, 665 (1995a) 
1684: % title: The jet-disk symbiosis I. Radio to X-ray emission models for
1685: % quasars
1686: %; astro-ph/9411096
1687: 
1688: \bibitem{1428} Falcke H., Biermann P.L., \textit{Astron. \& Astroph.} 
1689: \textbf{308}, 321 (1995b) 
1690: % title: Galactic jet sources and the AGN connection
1691: %; astro-ph/9506138
1692: 
1693: \bibitem{1432} Falcke H., Biermann P.L., \textit{Astron. \& Astroph.} 
1694: \textbf{342}, 49 - 56 (1999) 
1695: % title: The jet/disk symbiosis III. What the radio cores in
1696: % GRS1915+105, NGC4258, M81 and Sgr A$^{\star}$ tell us about
1697: % accreting black holes,
1698: %; astro-ph/9810226
1699: 
1700: \bibitem{1439} Falcke H., Malkan M.A., Biermann P.L., \textit{Astron. \&
1701: Astroph.} \textbf{298}, 375 (1995) 
1702: % title: The jet-disk symbiosis II. Interpreting the radio/UV correlations
1703: % in quasars
1704: %; astro-ph/9411100
1705: 
1706: \bibitem{1445} Falcke H., Sherwood W., Patnaik A.R., \textit{Astrophys. J.} 
1707: \textbf{471}, 106 (1996) 
1708: % title: The Nature of Radio-intermediate Quasars: What Is Radio-loud and 
1709: %  What Is Radio-quiet?
1710: %; astro-ph/9605165
1711: 
1712: \bibitem{1451} Faye G., Blanchet L., Buonanno A., \textit{Phys. Rev.} 
1713: \textbf{D} \textbf{74}, 104033 (2006).
1714: 
1715: \bibitem{1454} %  BBH-f-0612139v1
1716: Ferrarese L., Cote P., Blakeslee J.P., Mei S., Merritt D., West M.J., in 
1717: \textit{IAU Sympos.}, \textbf{238}, in press (2006a); astro-ph/0612139 
1718: %  title:  The inner workings of early type galaxies.  Cores, nuclei and 
1719: %  supermassive black holes
1720: %  reviews... as revealed by the ACS survey
1721: 
1722: \bibitem{1461} Ferrarese L. et al., \textit{Astrophys. J.} \ \textit{Suppl.}%
1723: \ \textbf{164}, 334 (2006b)
1724: 
1725: \bibitem{1464} Flanagan E.E., Hinderer T., \textit{Phys. Rev. D} \textbf{75}%
1726: , 124007 (2007) %[arXiv:0704.0389 [gr-qc]]
1727: 
1728: \bibitem{1467} Gergely L.\'{A}., Perj\'{e}s Z.. Vas\'{u}th M., \textit{Phys.
1729: Rev.} \textbf{D} \textbf{57}, 876 (1998a)
1730: 
1731: \bibitem{1470} Gergely L.\'{A}., Perj\'{e}s Z.. Vas\'{u}th M., \textit{Phys.
1732: Rev.} \textbf{D} \textbf{57}, 3423 (1998b)
1733: 
1734: \bibitem{1473} Gergely L.\'{A}., Perj\'{e}s Z.. Vas\'{u}th M., \textit{Phys.
1735: Rev.} \textbf{D} \textbf{58}, 124001 (1998)
1736: 
1737: \bibitem{1476} Gergely L.\'{A}., \textit{Phys. Rev.} \textbf{D} \textbf{61},
1738: 024035 (2000a)
1739: 
1740: \bibitem{1479} Gergely L.\'{A}., \textit{Phys. Rev.} \textbf{D} \textbf{62},
1741: 024007 (2000b)
1742: 
1743: \bibitem{1482} Gergely L.\'{A}., Keresztes Z., \textit{Phys. Rev.} \textbf{D}
1744: \textbf{67}, 024020 (2003)
1745: 
1746: \bibitem{} Gergely L.\'{A}., Mik\'{o}czi B., \textit{Phys. Rev. }\textbf{D} 
1747: \textbf{79}, 064023 (2009) %arXiv:0808.1704v2 [gr-qc] (2008)
1748: % title: "Renormalized 2PN spin contributions to the accumulated orbital phase for LISA sources"; 
1749: 
1750: \bibitem{1485} Ghez A. M., Salim S., Hornstein S.D., et al., \textit{%
1751: Astrophys. J.} \textbf{620}, 744 - 757 (2005) 
1752: % title: Stellar Orbits around the Galactic Center Black Hole
1753: % complete authior list: Ghez, A. M., Salim, S., Hornstein, S. D., Tanner, A., 
1754: % Lu, J. R., Morris, M., Becklin, E. E., Duchne, G.,
1755: % Result: The estimated central dark mass from orbital motions is 3.7(+/-0.2)
1756: % 10^{6}[R0/(8kpc)]^{3} Msolar
1757: %; astro-ph/0306130
1758: 
1759: \bibitem{1494} Gilmore G., Wilkinson M., Kleyna J., Koch A., Wyn Evans N.,
1760: Wyse R.F.G., Grebel E.K., presented at UCLA Dark Matter 2006 Conference,
1761: March 2006, \textit{Nucl. Phys. Proc. Suppl.} \textbf{173}, 15 (2007) 
1762: % title: Observed Properties of Dark Matter: dynamical studies of dSph
1763: % galaxies
1764: % -- has the famous graph M/L versus L for dwarf spheroidals, which shows that 
1765: % they all are limited by M_{DM} > 5 10^(7) solar masses
1766: %astro-ph/0608528
1767: 
1768: \bibitem{1502} Gonz{\'{a}}lez J.A. et al. \textit{Phys. Rev. Letters} 
1769: \textbf{98}, 091101 (2007a) 
1770: %  title:  Maximum Kick from Nonspinning Black-Hole Binary Inspiral
1771: %  When unequal-mass black holes merge, the final black hole receives a kick due %  to the asymmetric loss of linear momentum in the gravitational radiation 
1772: %  emitted during the merger. The magnitude of this kick has important 
1773: %  astrophysical consequences. Recent breakthroughs in numerical relativity 
1774: %  allow us to perform the largest parameter study undertaken to date in 
1775: %  numerical simulations of binary black-hole inspirals. We study nonspinning 
1776: %  black-hole binaries with mass ratios from q=M1/M2=1 to q=0.25 (?=q/(1+q)2 
1777: %  from 0.25 to 0.16). We accurately calculate the velocity of the kick to 
1778: %  within 6 percent, and the final spin of the black holes to within 2 percent . %  A maximum kick of 175.211kms-1 is achieved for ?=0.1950.005
1779: 
1780: \bibitem{1514} Gonz{\'{a}}lez J.A., Hannam M.D., Sperhake U., Br{\"{u}}gmann
1781: B., Husa S., \textit{Phys. Rev. Lett.} \textbf{98}, 231101 (2007b) 
1782: % title: Supermassive kicks for spinning black holes
1783: % ... kick velocities of at least $2500 $km/s are possible for equal-mass
1784: % binaries with appropriate alignments of the spins. Strong for equal masses.
1785: %gr-qc/0702052
1786: 
1787: \bibitem{1520} Gopal-Krishna, Wiita P.J., \textit{Astrophys. J.} \textbf{529}%
1788: , 189 - 200 (2000) % title: Superdisks in Radio Galaxies
1789: 
1790: \bibitem{1523} Gopal-Krishna, Biermann P.L., Wiita P.J., \textit{Astrophys.
1791: J. Letters} \textbf{594}, L103 - L106 (2003) 
1792: % title: The Origin of the X-shaped radio galaxies: Clues from the
1793: % Z-symmetric Secondary Lobes
1794: %; astro-ph/0308059
1795: 
1796: \bibitem{1529} Gopal-Krishna, Biermann P.L., Wiita P.J., \textit{Astrophys.
1797: J. Letters} \textbf{603}, L9 - L12 (2004) 
1798: % title: Brightness suppression of relativistic radio jets of quasars: The
1799: % role of the lower electron cut-off
1800: %; astro-ph/0401463
1801: 
1802: \bibitem{1535} Gopal-Krishna, Wiita P.J., Joshi S., \textit{Month. Not. Roy.
1803: Astr. Soc.} \textbf{380}, 703 (2007) 
1804: % title: Superdisks in radio galaxies: Jet-wind interactions
1805: 
1806: \bibitem{1539} Gopal-Krishna, Wiita P.J., invited talk at the 4th Korean
1807: Workshop on high energy astrophysics (April 2006),
1808: http://sirius.cnu.ac.kr/kaw4/presentations.htm % title:
1809: 
1810: \bibitem{1543} Gopal-Krishna, Zier Ch., Gergely L.\'{A}, Biermann P.L.,
1811: (2009), in preparation
1812: 
1813: \bibitem{1546} Gott III J.R., Turner E.L., \textit{Astrophys. J.} \textbf{216%
1814: }, 357 (1977)
1815: 
1816: \bibitem{1549} H\"{a}ring N., Rix H., \textit{Astrophys. J. Letters} , 
1817: \textbf{604}, L89 (2004)
1818: 
1819: \bibitem{1552} Hayasaki K., to appear in \textit{Publications of the
1820: Astronomical Society of Japan}, arXiv:0805.3408 (2008) 
1821: %title: A new mechanism for massive binary black hole evolution
1822: 
1823: \bibitem{1556} Hickson P., \textit{Astrophys. J.} , \textbf{255}, 382 (1982)
1824: 
1825: \bibitem{1558} Hughes S.A., Blandford R.D., \textit{Astrophys. J.} \textbf{%
1826: 585}, L101 (2003)
1827: 
1828: \bibitem{1561} Ioka K., Taniguchi K., \textit{Astrophys. J.} \textbf{537},
1829: 327 - 333 (2000) 
1830: %  title:  Gravitational Waves from Inspiraling Compact Binaries with Magnetic 
1831: %  Dipole Moments
1832: 
1833: \bibitem{1566} Iye M. et al. \textit{Nature} \textbf{443}, 186 - 188 (2006) 
1834: % title:  A galaxy at a redshift z = 6.96
1835: %  When galaxy formation started in the history of the Universe remains unclear. %  Studies of the cosmic microwave background indicate that the Universe, after %  initial cooling (following the Big Bang), was reheated and reionized by hot 
1836: %  stars in newborn galaxies at a redshift in the range 6 < z < 14 (ref. 1). 
1837: %  Though several candidate galaxies at redshift z > 7 have been identified 
1838: %  photometrically, galaxies with spectroscopically confirmed redshifts have 
1839: %  been confined to z < 6.6 (refs 4-8). Here we report a spectroscopic redshift %  of z = 6.96 (corresponding to just 750Myr after the Big Bang) for a galaxy 
1840: %  whose spectrum clearly shows Lyman-? emission at 9,682, indicating active 
1841: %  star formation at a rate of ~10Msolaryr-1, where Msolar is the mass of the 
1842: %  Sun. This demonstrates that galaxy formation was under way when the Universe %  was only ~6 per cent of its present age. The number density of galaxies at z %  ~ 7 seems to be only 18-36 per cent of the density at z = 6.6.
1843: 
1844: \bibitem{1577} Junor W., Salter C.J., Saikia D.J., Mantovani F., Peck A.B., 
1845: \textit{Month. Not. Roy. Astr. Soc.} \textbf{308}, 955 - 960 (1999) 
1846: % title: Large differential Faraday rotation in the compact steep-spectrum
1847: % quasar 3C 147 and jet-medium interactions.. . We suggest a consistent 
1848: %  picture for 3C 147, and comment on the evolution of compact steep-spectrum 
1849: %  radio sources in general.
1850: 
1851: \bibitem{1584} Kidder L., Will C., Wiseman A., \textit{Phys. Rev.} \textbf{D}
1852: \textbf{47}, R4183 (1993)
1853: 
1854: \bibitem{1587} Kidder L., \textit{Phys. Rev.} \textbf{D} \textbf{52}, 821
1855: (1995)
1856: 
1857: \bibitem{1589} King A.R., Pringle J.E., \textit{Month. Not. Roy. Astr. Soc.} 
1858: \textbf{373}, L90 (2006)
1859: 
1860: \bibitem{1592} Klypin A., Zhao H.-S., Somerville R.S., \textit{Astrophys. J.}
1861: \textbf{573}, 597 - 613 (2002) 
1862: % title: CDM-based Models for the Milky Way and M31. I. Dynamical Models
1863: %; astro-ph/0110390
1864: 
1865: \bibitem{} Koppitz, M., Pollney, D., Reisswig, C., Rezzolla, L., Thornburg,
1866: J., Diener, P., Schnetter E., \textit{Phys. Rev. Lett.} \textbf{99}, 041102
1867: (2007) %arXiv:gr-qc/0701163
1868: %Title: Recoil Velocities from Equal-Mass Binary-Black-Hole Mergers
1869: 
1870: \bibitem{1597} Kormendy J., Richstone D., \textit{Annual Rev. of Astron. \&
1871: Astrophys.} \textbf{33}, 581 (1995) 
1872: % title: Inward Bound---The Search For Supermassive Black Holes In Galactic
1873: % Nuclei
1874: 
1875: \bibitem{} Kov\'{a}cs Z., Biermann, P.L., Gergely, \'{A}.L., "The maximal
1876: spin of a black hole, disk and jet symbiotic system" in preparation (2009)
1877: 
1878: \bibitem{} Lang R.N., Hughes S.A., \textit{Phys. Rev.} \textbf{D} \textbf{74}%
1879: , 122001 (2006). Errata, ibid. \textbf{D} \textbf{75}, 089902(E) (2007)
1880: 
1881: \bibitem{} Lang R.N., Hughes S.A., \textit{Astrophys. J.} \textbf{677}, 1184
1882: (2008)
1883: 
1884: \bibitem{1602} Lauer T.R. et al., \textit{Astrophys. J.} \textbf{662}, 808
1885: (2007)%; astro-ph/0606739
1886: 
1887: \bibitem{1605} Lawrence A., Elvis M. \textit{Astrophys. J.} \textbf{256},
1888: 410 - 426 (1982) 
1889: %  title:  Obscuration and the various kinds of Seyfert galaxies
1890: % on torus in AGN
1891: 
1892: \bibitem{1609} Lousto C.O., Zlochower Y., \textit{Phys. Rev.} \textbf{D 79},
1893: 064018 (2009) 
1894: %  title:  "Modeling gravitational recoil from precessing highly-spinning unequal-mass black-hole binaries"; 
1895: %arXiv:0805.0159v2 [gr-qc] (2008)
1896: 
1897: \bibitem{1613} Lynden-Bell D., \textit{Month. Not. Roy. Astr. Soc.} \textbf{%
1898: 136}, 101 (1967) 
1899: %  title: Statistical mechanics of violent relaxation in stellar systems
1900: % invention of what is now referred to as "violent relaxation"
1901: 
1902: \bibitem{1618} Mahadevan R., \textit{Nature} \textbf{394}, 651 - 653 (1998) 
1903: % title: Reconciling the spectrum of Sagittarius A* with a two-temperature
1904: % plasma model.
1905: 
1906: \bibitem{1622} %  BBH-b-58642.web
1907: Makino J., Funato Y., \textit{Astrophys. J.} \textbf{602}, 93 - 102 (2004) 
1908: %  title:  Evolution of massive black hole binaries
1909: %  loss cone arguments; conclusion black hole binaries do not merge
1910: 
1911: \bibitem{1627} Marcaide J.M., Shapiro I.I., \textit{Astron. J.} \textbf{88},
1912: 1133 - 1137 (1983) 
1913: %  title:  High precision astrometry via very-long-baseline radio interferometry %  - Estimate of the angular separation between the quasars 1038+528A and B
1914: % errors in the relative distance measurements of micro arc seconds
1915: 
1916: \bibitem{1632} Marecki A., Barthel P. D., Polatidis A., Owsianik I., \textit{%
1917: Publ. Astron. Soc. Australia} \textbf{20}, 16 - 18 (2003) 
1918: % title: 1245+676 - A CSO/GPS Source being an Extreme Case of a Double-Double 
1919: % Structure
1920: 
1921: \bibitem{1637} %  BBH-a-64183.web
1922: Matsubashi T., Makino J., Ebisuzaki T.\textit{Astrophys. J.} \textbf{656},
1923: 879 - 896 (2007) 
1924: %  title:  Orbital evolution of an intermediate mass black hole in the galactic 
1925: %  nucleus with a massive black hole
1926: %  100 - 10^{5} solar masses
1927: 
1928: \bibitem{1644} Matsui H., Habe A., Saitoh T.R., \textit{Astrophys. J.} 
1929: \textbf{651}, 767 - 774 (2006) 
1930: % title: Effects of a Supermassive Black Hole Binary on a Nuclear Gas Disk
1931: % We study the influence of a galactic central supermassive black hole (SMBH)
1932: % binary on gas dynamics and star formation activity in a nuclear gas disk 
1933: % three-dimensional tree+SPH simulations. .. small starbursts, resonances 
1934: %; astro-ph/0606140
1935: 
1936: \bibitem{1652} %  BBH-j-0301257v1
1937: Merritt D., in Proc. \textit{Coevolution of black holes and galaxies},
1938: Cambridge U. Press, Ed. L.C. Ho (in press) (2003), astro-ph/0301257 
1939: %  title:  Single and binary black holes and the influence on nuclear structure
1940: 
1941: \bibitem{1657} %  BBH-i-19064.web
1942: Merritt D. \textit{Astrophys. J. Letters} \textbf{621}, L101 - L104 (2005) 
1943: %  title:  Loss cone refilling rates in galactic nuclei
1944: %  very long loss cone refilling times are derived
1945: 
1946: \bibitem{1662} %  BBH-m-1310
1947: Merritt D. \& Ekers R., \textit{Science} \textbf{297}, 1310-1313 (2002) 
1948: % title: Tracing Black Hole Mergers Through Radio Lobe Morphology
1949: %; astro-ph/0208001
1950: 
1951: \bibitem{1667} %  BH-c-0705.2745v1
1952: Merritt D., Mikkola S., Szell A., arXiv/0705.2745 %  DRAFT version!!!!
1953: %  title:  Long term evolution of massive black hole binaries III: Binary 
1954: %  evolution in collisional nuclei
1955: %  prevent stalling by interaction with stars
1956: %  does quote three papers Zier +, Merritt et al agree with Zier...
1957: 
1958: \bibitem{1676} Merritt D., Miloslavljevi\'{c} M., \textit{Living Rev.
1959: Relativity} \textbf{8}, 8 (2005)
1960: 
1961: \bibitem{1679} Mik\'{o}czi B, Vas\'{u}th M, Gergely L.\'{A}., \textit{Phys.
1962: Rev.} \textbf{D} \textbf{71}, 124043 (2005)
1963: 
1964: \bibitem{1682} %  BBH-l-0212270.v2
1965: Milosavljevi{\'{c}} M., Merritt D., "The Final Parsec Problem " , \textit{%
1966: AIP Proc.} (in press), (2003a); astro-ph/0212270 
1967: %  title:  The final parsec problem
1968: %  mostly about the loss cone problem
1969: 
1970: \bibitem{1687} %  BBH-k-57515.web
1971: Milosavljevi{\'{c}} M., Merritt D., \textit{Astrophys. J.} \textbf{596}, 860
1972: - 878 (2003b) %  title:  Long term evolution of masive black hole binaries
1973: %  the final parsec problem,  various regimes are discussed..
1974: %  emphasizes the Brownian motion
1975: %  mostly about the loss cone problem
1976: 
1977: \bibitem{1694} Munyaneza F., Biermann P.L., \textit{Astron. \& Astroph.} 
1978: \textbf{436}, 805 - 815 (2005) 
1979: % title: Fast Growth of supermassive black holes in Galaxies, ; astro-ph/0403511
1980: 
1981: \bibitem{1698} Munyaneza F., Biermann P.L., \textit{Astron. \& Astroph.
1982: Letters} \textbf{458}, L9 - L12 (2006) 
1983: % title: Degenerate sterile neutrino dark matters in the cores of galaxies, ; astro-ph/0609388
1984: 
1985: \bibitem{1702} Mushotzky, R. \textit{Astrophys. J.} \textbf{256}, 92 - 102
1986: (1982) 
1987: %  title:  The X-ray spectrum and time variability of narrow emission line 
1988: %  galaxies
1989: % ---  torus in X-ray absorption, in AGN
1990: 
1991: \bibitem{1708} Nagar N.M., Falcke H., Wilson A.S., Ho L.C., \textit{%
1992: Astrophys. J.} \textbf{542}, 186 - 196 (2000) 
1993: % title: Radio Sources in Low-Luminosity Active Galactic Nuclei. I. VLA
1994: % Detections of Compact, Flat-Spectrum Cores, ; astro-ph/0005382
1995: 
1996: \bibitem{1713} O'Connell R.F., \textit{Phys. Rev. Letters} \textbf{93},
1997: 081103 (2004) 
1998: % title: Proposed New Test of Spin Effects in General Relativity 
1999: 
2000: \bibitem{1716} Owen F.N., Eilek J.A., Kassim N.E., \textit{Astrophys. J.} 
2001: \textbf{543}, 611 (2000) 
2002: % title: M87 at 90 Centimeters: A Different Picture
2003: 
2004: \bibitem{1720} Perez-Fournon I., Biermann P.L., \textit{Astron. \& Astroph.
2005: Letters} \textbf{130}, L13 - L15 (1984) 
2006: % title: Do all bright elliptical galaxies have active nuclei?
2007: 
2008: \bibitem{1724} Peters P.C., \textit{Phys. Rev.} \textbf{136}, B1224 (1964)
2009: 
2010: \bibitem{1726} Peters P.C., Mathews S., \textit{Phys. Rev.} \textbf{131},
2011: 435\ (1963)
2012: 
2013: \bibitem{1729} Poisson E., \textit{Phys. Rev.} \textbf{D} \textbf{57}, 5287
2014: (1998)
2015: 
2016: \bibitem{1732} Press W.H., Schechter P., \textit{Astrophys. J.} , \textbf{187%
2017: }, 425 (1974)
2018: 
2019: \bibitem{1735} Pretorius F., in \textit{Relativistic Objects in Compact
2020: Binaries: From Birth to Coalescence}, ed. Colpi et al., Springer Verlag,
2021: Canopus Publishing Limited, arXiv:0710.1338 [gr-qc] (2007) 
2022: % title: Binary Black Hole Coalescence
2023: %arXiv:0803.1820 [gr-qc] 
2024: \ 
2025: 
2026: \bibitem{1737} Racine E., \textit{Phys. Rev.} \textbf{D 78}, 044021 (2008)
2027: 
2028: \bibitem{} Rezzolla L., Barausse E., Dorband E. N., Pollney D., Reisswig
2029: Ch., Seiler J. , Husa S., \textit{Phys. Rev.} \textbf{D 78}, 044002 (2008a) 
2030: %On the final spin from the coalescence of two black holes, arXiv:0712.3541
2031: 
2032: \bibitem{} Rezzolla L., Diener P., Dorband E.N., Pollney D., Reisswig Ch.,
2033: Schnetter E., Seiler J., \textit{Astrophys. J.} \textbf{674}, L29 (2008b) 
2034: % The final spin from the coalescence of aligned-spin black-hole binaries, arXiv:0710.3345 
2035: 
2036: \bibitem{} Rezzolla L, Dorband E.N., Reisswig Ch., Diener P., Pollney D.,
2037: Schnetter E., Szilagyi B., \textit{Astrophys. J.} \textbf{679}, 1422 (2008c) 
2038: %  Spin Diagrams for Equal-Mass Black-Hole Binaries with Aligned Spins, arXiv:0708.3999 
2039: 
2040: \bibitem{1741} Richter O.-G., Sackett P.D., Sparke L.S., \textit{Astron. J.} 
2041: \textbf{107}, 99 - 117 (1994) 
2042: % title: A neutral hydrogen survey of polar-ring galaxies, 1: Green Bank
2043: % observations of the northern sample: includes NGC2685
2044: %; astro-ph/9308023
2045: 
2046: \bibitem{1746} Rieth R., Sch\"afer G., \textit{Class. Quantum Grav.} \textbf{%
2047: 14}, 2357 (1997)
2048: 
2049: \bibitem{1749} Roman S.-A., Biermann P.L., \textit{Roman. Astron J. Suppl.}, 
2050: \textbf{16}, 147 (2006)
2051: 
2052: \bibitem{1752} Rottmann H., PhD thesis: "Jet-Reorientation in X-shaped Radio
2053: Galaxies", Bonn Univ., 2001: (http://hss.ulb.uni-bonn.de/diss$\_$online/math 
2054: $\_$nat$\_$fak/2001/rottmann$\_$helge/index.htm) 
2055: % title: Jet-Reorientation in X-shaped Radio Galaxies
2056: 
2057: \bibitem{1757} Ryan F., \textit{Phys. Rev.} \textbf{D 53}, 3064 (1996)
2058: 
2059: \bibitem{1759} Sanders D.B., Mirabel I.F., \textit{Annual Rev. of Astron. \&
2060: Astrophys.} \textbf{34}, 749 (1996) % title: Luminous Infrared Galaxies
2061: 
2062: \bibitem{1762} Sch{\"{a}}fer, G., \textit{Current Trends in Relativistic
2063: Astrophysics}, Edited by L. Fern{\'{a}}ndez-Jambrina, L.M. Gonz{\'{a}}%
2064: lez-Romero, Lecture Notes in Physics, vol. \textbf{617}, p. 195 (2005) 
2065: %  title:  Binary Black Holes and Gravitational Wave Production: Post-Newtonian %  Analytic Treatment
2066: %  Within a post-Newtonian approximation scheme the Einstein field equations are %  solved for the dynamics of the decaying orbits of binary black holes and the %  related gravitational wave emission. The black holes are modelled by two 
2067: %  Dirac delta functions in conformally related d-dimensional euclidean space. 
2068: %  The limit from the Einstein field equations in (d+1)-dimensional spacetime is %  applied to achieve well-defined mathematical expressions in 4-dimensional 
2069: %  spacetime. The conservative orbital dynamics is presented up to the third 
2070: %  post-Newtonian order of approximation and the decaying orbital phase up to 
2071: %  the four-and-a-half post-Newtonian order. The gravitational waveform is given %  to second post-Newtonian order.
2072: 
2073: \bibitem{1774} Sch{\"{o}}del R., Eckart A., \textit{Mem. Soc. Astron. Ital.} 
2074: \textbf{76}, 65 (2005) 
2075: % title: The Centre of the Milky Way: Stellar Dynamics, Potential Star
2076: % Formation, and Variable NIR Emission from Sgr A*
2077: 
2078: \bibitem{1779} Sesana A., Haardt F., Madau P., \textit{Astrophys. J.} 
2079: \textbf{651}, 392S (2006) 
2080: % title: Interaction of massive black hole binaries with their stellar environment: I. Ejection of Hypervelocity Stars
2081: 
2082: \bibitem{1783} Sesana A., Haardt F., Madau P., \textit{Astrophys. J.} 
2083: \textbf{660}, 546S (2007a) 
2084: % title: Interaction of massive black hole binaries with their stellar environment: II. Loss Cone Depletion and Binary Orbital Decay
2085: 
2086: \bibitem{1787} Sesana A., Haardt F., Madau P., to appear in \textit{%
2087: Astrophys. J.}; arXiv:0710.4301 (2007b) 
2088: % title: Interaction of massive black hole binaries with their stellar environment: III. Scattering of bound stars 
2089: 
2090: \bibitem{1791} Silk J., Takahashi T., \textit{Astrophys. J.} \textbf{229},
2091: 242 - 256 (1979) 
2092: % title: A statistical model for the initial stellar mass function
2093: 
2094: \bibitem{1795} Thorne, K. S., \textit{Proc. Royal Soc. London} \textbf{A 368}%
2095: , 9 (1979) % Sources of Gravitational Waves
2096: 
2097: \bibitem{1798} Toomre A., Toomre J., \textit{Astrophys. J.} \textbf{178},
2098: 623 - 666 (1972) % title: Galactic Bridges and Tails
2099: 
2100: \bibitem{1801} %  BBH-t-1996MNRAS_278_186
2101: Valtonen M.J., \textit{Month. Not. Roy. Astr. Soc.} 278, 186 (1996) 
2102: %  Triple black hole systems formed in mergers of galaxies
2103: %  We study the evolution of orbits of supermassive black holes in the nuclei of 
2104: %  galaxies under the gravitational influence of the stars in the galaxy, as 
2105: %  well as the cooling flow.... we find that successive mergers create few black 
2106: %  hole systems..
2107: 
2108: \bibitem{1809} Vas\'{u}th M, Keresztes Z, Mih\'{a}ly A., Gergely L .\'{A}., 
2109: \textit{Phys. Rev.} \textbf{D} \textbf{68}, 124006 (2003)
2110: 
2111: \bibitem{1812} Volonteri M., Madau P., Quataert E., Rees M. J., \textit{%
2112: Astrophys. J.} \textbf{620}, 69 (2005)
2113: 
2114: \bibitem{1815} Wilson A.S., Colbert E.J.M., \textit{Astrophys. J.} \textbf{%
2115: 438}, 62 - 71 (1995)
2116: 
2117: \bibitem{1818} Worrall D.M., Birkinshaw M., Kraft R.P., Hardcastle M.J., 
2118: \textit{Astrophys. J. Letters}, in press (2007); astro-ph/0702411 
2119: % title: The effect of a Chandra-measured merger-related gas component on the
2120: % lobes of a dead radio galaxy
2121: 
2122: \bibitem{1823} %  BBH-w-q31007
2123: Yu Q., \textit{Class. Quantum Grav.} \textbf{20}, S55-S63 (2003) 
2124: %  title:  Evolution of massive binary black holes
2125: %  claims that many BBH survive as binaries
2126: 
2127: \bibitem{1828} Zier Ch., Biermann P.L., \textit{Astron. \& Astroph.} \textbf{%
2128: 377}, 23 - 43 (2001) 
2129: % title: Binary Black Holes and Tori in AGN I. Ejection of stars and
2130: % merging of the binary
2131: %; astro-ph/0106419
2132: 
2133: \bibitem{1833} Zier Ch., Biermann P.L., \textit{Astron. \& Astroph.} \textbf{%
2134: 396}, 91 (2002) 
2135: % title: Binary Black Holes and tori in AGN II. Can stellar winds
2136: % constitute a dusty torus?
2137: %; astro-ph/0203359
2138: 
2139: \bibitem{1838} Zier Ch., \textit{Month. Not. Roy. Astr. Soc.} \textbf{364},
2140: 583 (2005) 
2141: % title: Orientation and size of the `Z' in X-shaped radio galaxies
2142: %; astro-ph/0507129
2143: 
2144: \bibitem{1842} Zier Ch., \textit{Month. Not. Roy. Astr. Soc.} \textit{Lett.} 
2145: \textbf{371}, L36 - L40 (2006) % paper ChZ-b
2146: % title: Merging of a massive binary due to ejection of bound stars
2147: %  Merging of a massive binary due to ejection of bound stars
2148: %  Previous arguments centered on the stalling of the shrinking process due to 
2149: %  loss-cone depletion:  Here we follow a different approach, and focus on the 
2150: %  population of stars which is bound to the binary.  With simple assumptions we 
2151: %  derive a lower limit for the mass of stars which needs to be ejected by the 
2152: %  binary in order to coalesce.  Our results suggest that binary BH merge.
2153: %; astro-ph/0605619
2154: 
2155: \bibitem{1852} Zier Ch., \textit{Month. Not. Roy. Astr. Soc.} \textbf{378},
2156: 1309-1327 (2007) % title: Merging of a massive black hole binary II
2157: %; astro-ph/0610457
2158: 
2159: \bibitem{1855} Zier Ch., Gergely L.\'{A}., Biermann P.L., (2009), in
2160: preparation
2161: \end{thebibliography}
2162: 
2163: \end{document}
2164: