gr-qc0508067/bcd.tex
1: \documentclass[prd,aps,showpacs,floats,letterpaper,floatfix,groupedaddress,eqsecnum,nofootinbib]{revtex4}
2: 
3: \usepackage{amssymb,amsmath}
4: \usepackage[dvips]{graphicx}
5: 
6: \special{papersize=8.5in,11in}
7: 
8: \usepackage{longtable}
9: \usepackage{dcolumn,epsfig}
10: 
11: \def\np{({\bf n}\cdot{\bf p})}
12: \def\pp{{\bf p}^2}
13: \def\ppp{({\bf p}^2)}
14: 
15: \def\laq{\raise 0.4ex\hbox{$<$}\kern -0.8em\lower 0.62ex\hbox{$\sim$}}
16: \def\gaq{\raise 0.4ex\hbox{$>$}\kern -0.7em\lower 0.62ex\hbox{$\sim$}}
17: 
18: \newlength{\sizeonefig}
19: \newlength{\sizetwofig}
20: \setlength{\sizeonefig}{0.48\textwidth}
21: \setlength{\sizetwofig}{0.48\textwidth}
22: 
23: \newcommand{\vphi}{\varphi} 
24: \newcommand{\pa}{\partial} 
25: \newcommand{\hL}{\mbox{\boldmath${\hat{L}}$}}
26: \newcommand{\hS}{\mbox{\boldmath${\hat{S}}$}}
27: \newcommand{\hP}{\mbox{\boldmath${\hat{P}}$}}
28: \newcommand{\hl}{\mbox{\boldmath${\hat{\lambda}}$}}
29: \newcommand{\hn}{\mbox{\boldmath${\hat{n}}$}}
30: \newcommand{\vx}{\mbox{\boldmath${x}$}}
31: \newcommand{\vp}{\mbox{\boldmath${p}$}}
32: \newcommand{\vq}{\mbox{\boldmath${q}$}}
33: \newcommand{\vv}{\mbox{\boldmath${v}$}}
34: \newcommand{\vV}{\mbox{\boldmath${V}$}}
35: \newcommand{\va}{\mbox{\boldmath${a}$}}
36: \newcommand{\vJ}{\mbox{\boldmath${J}$}}
37: \newcommand{\vF}{\mbox{\boldmath${F}$}}
38: \newcommand{\vA}{\mbox{\boldmath${A}$}}
39: \newcommand{\vO}{\mbox{\boldmath${\Omega}$}}
40: \newcommand{\vP}{\mbox{\boldmath${P}$}}
41: \newcommand{\vD}{\mbox{\boldmath${\Delta}$}}
42: \newcommand{\vS}{\mbox{\boldmath${S}$}}
43: \newcommand{\vN}{\mbox{\boldmath${N}$}}
44: \newcommand{\vX}{\mbox{\boldmath${X}$}}
45: \newcommand{\vL}{\mbox{\boldmath${L}$}}
46: \newcommand{\vlb}{\mbox{\boldmath${\lambda}$}}
47: \newcommand{\vs}{\mbox{\boldmath${s}$}}
48: \newcommand{\beq}{\begin{equation}}
49: \newcommand{\eeq}{\end{equation}}
50: \newcommand{\bea}{\begin{eqnarray}} 
51: \newcommand{\eea}{\end{eqnarray}}
52: \newcommand{\ba}{\begin{array}}
53: \newcommand{\ea}{\end{array}}
54: 
55: \newcommand{\comment}[1]{}
56: 
57: \begin{document}
58: 
59: \title{Transition from inspiral to plunge in precessing binaries of spinning black holes}
60: 
61: \author{Alessandra Buonanno}
62: 
63: \affiliation{AstroParticule et Cosmologie (APC),
64: 11, place Marcelin Berthelot, 75005 Paris, France}
65: 
66: \altaffiliation{UMR 7164 (CNRS, Universit\'e Paris7, CEA, Observatoire
67: de Paris). Also: Institut d'Astrophysique de Paris, 98$^{\rm bis}$ 
68: Boulv. Arago, 75013 Paris, France.}
69: 
70: \author{Yanbei Chen}
71: 
72: \affiliation{Max-Planck-Institut f\"ur Gravitationsphysik
73: (Albert-Einstein-Institut), Am M\"uhlenberg 1, D-14476 Golm bei
74: Potsdam, Germany}
75: 
76: \author{Thibault Damour}
77: 
78: \affiliation{Institut des Hautes Etudes Scientifiques, 91440 Bures-sur-Yvette, France}
79: 
80: \begin{abstract}
81: We investigate the non-adiabatic dynamics of spinning black hole
82: binaries by using an analytical Hamiltonian completed with a radiation-reaction force,
83: containing spin couplings, which matches the known rates of energy and
84: angular momentum losses on quasi-circular orbits. We consider
85: both a straightforward post-Newtonian-expanded Hamiltonian
86: (including spin-dependent terms), and a version of
87: the resummed post-Newtonian
88: Hamiltonian defined by the Effective One-Body approach.
89: We focus on the influence of spin terms onto the dynamics and waveforms.
90: We evaluate the energy and angular momentum released during the final
91: stage of inspiral and plunge. For an equal-mass
92: binary the energy released between 40\,Hz and the frequency beyond
93: which our analytical treatment becomes unreliable is found to be,
94: when using the more reliable Effective One-Body dynamics:
95:  $0.6\% M$ for  anti-aligned maximally spinning black holes,
96:  $5\% M$ for aligned maximally spinning black hole, and
97: $1.8\% M$ for non-spinning configurations. 
98: In confirmation of previous results, we find that, for all binaries considered,
99: the dimensionless rotation
100: parameter $J/E^2$ is always smaller than unity at the end of the inspiral, 
101: so that a Kerr black hole can form right after the inspiral phase.
102: By matching a quasi-normal mode ringdown to the last reliable stages
103: of the plunge, we construct complete waveforms approximately describing the
104: gravitational wave signal emitted by the entire process of coalescence
105: of precessing binaries of spinning black holes.
106: \end{abstract}
107: 
108: \maketitle
109: 
110: \section{Introduction}
111: \label{sec1}
112: 
113: An international network of kilometer-scale laser-interferometric
114: gravitational-wave detectors, consisting of the Laser-Interferometer
115: Gravitational-wave Observatory (LIGO) \cite{LIGO}, of VIRGO
116: \cite{VIRGO}, of GEO\,600 \cite{GEO} and of TAMA\,300 \cite{TAMA}, has
117: by now begun the science operations. TAMA\,300 reached its full
118: design sensitivity in 2001, VIRGO is in its commissioning phase
119: and plans to start the first scientific runs by the end of 2005, 
120: while LIGO has already completed three science runs (two of them 
121: in coincidence with GEO\,600) with increasing sensitivity and duty 
122: cycle. LIGO and GEO\,600 are expected to reach their full design 
123: sensitivity in 2005.
124: 
125: Binary black holes are among the most promising sources for these
126: detectors. Among black hole binaries, it was emphasized in \cite{TD} that
127: there is a bias towards first detecting mostly aligned spinning
128: binaries with high masses, whose last stable spherical orbits are drawn, 
129: by spin effects, to larger binding energies, yet due to their high masses 
130: these energies are still emitted through gravitational waves in the sensitive 
131: band of the detectors. Studying in detail the waveforms emitted during the 
132: last stages of dynamical evolution of such heavy spinning black hole binaries,
133: with explicit consideration of the crucial transition between
134: adiabatic inspiral and plunge, is a demanding theoretical challenge.
135: The aim of the present paper is to provide a first attack on this problem
136: by using some of the best analytical tools currently available to describe 
137: the transition from adiabatic inspiral to plunge, and notably
138: the Effective One Body (EOB) approach \cite{BD1,BD2}.
139: 
140: 
141: So far, most theoretical and data-analysis studies on precessing binaries of spinning black holes
142: assumed {\it adiabatic evolution}. Thus they were restricted to considering {\it only} 
143: the inspiral phase~\cite{ACST94,K,apostolatos0,apostolatos1,apostolatos2,
144: GKV,GK,bcv2,pbcv1,bcpv1,Gpc}. Actually, even for non-spinning
145: binary configurations, most theoretical studies confined themselves 
146: to considering the adiabatic inspiral phase, though a lot 
147: of effort was spent to improve the accuracy of the phasing during the last stages 
148: of the inspiral, see e.g. \cite{DIS98, DIS00}.
149: 
150: For heavy black hole binaries, most of the signal-to-noise ratio in the 
151: ground detectors will come from the very last stages of the
152: inspiral, and from the non-adiabatic transition between
153: inspiral and plunge. It is therefore essential to be able
154: to describe, with acceptable accuracy, this non-adiabatic evolution.
155: In Refs.~\cite{BD1,BD2} a new way of describing the dynamics of
156: binary systems was introduced: the Effective One-Body (EOB)
157: approach. The EOB approach uses both a specific resummation of the
158: post-Newtonian Hamiltonian, and a resummed version of
159: radiation reaction. This was shown to lead to a rather
160: robust formalism, which is likely to provide a reliable description
161: of non-adiabatic effects, of the transition between inspiral and
162: plunge, and of the beginning of the plunge. It was also used in
163: \cite{BD2} to model the full merger phase of
164: non-spinning binaries, by matching the natural end
165: of the EOB plunge with the ring-down phase. The EOB approach was used
166: in Refs.~\cite{BD2,DJS} to derive non-adiabatic template waveforms emitted
167: by {\it non-spinning} black hole binaries. It was shown in \cite{DIS01}
168: that these new EOB templates led to significantly enhanced signal to noise
169: ratios in current detectors (mainly because of the inclusion of
170: the plunge signal). The EOB Hamiltonian was extended to the
171: case of spinning black holes in \cite{TD}.
172: The analytical predictions made by the EOB method (including spin)
173: were found to agree remarkably well \cite{DGG} with the
174:  numerical results obtained by means of the helical Killing
175: vector approach \cite{GGB} for circular orbits of corotating
176: black holes. Several other studies showed that the EOB method
177: provides phasing models which are more reliable and robust
178: than other (adiabatic or non-adiabatic) models \cite{bcv1,DIS03}.
179: 
180: The main purpose of this paper is to extend the use of the EOB approach 
181: to the case of precessing binaries of spinning black holes,
182: both by including spin-dependent terms in radiation-reaction effects,
183: and by studying the waveforms generated beyond the adiabatic
184: approximation, i.e. taking into account the transition between
185: inspiral and plunge, and the plunge itself. Let us emphasize again
186: that the EOB approach has the advantage of providing an {\it analytical}
187: description of the transition from inspiral to plunge. Recently,
188: some attempts have been made to tackle, by means of 3D {\it numerical}
189: simulations (combined with a perturbative approach), the
190: gravitational radiation emitted by a very tight black hole binary both
191: in non-spinning~\cite{BBCLT}, and moderately spinning,
192: but non-precessing~\cite{BCLT} configurations. These three-dimensional (3D)
193: simulations concluded to the emission of a significantly larger amount of
194: energy in the form of gravitational radiation than what we shall
195: find from our analytical, EOB approach. It should be mentioned in
196: this respect that the energy released in the form of gravitational
197: waves depends very much on the choice of initial data, and that the
198: amount by which the initial data chosen in \cite{BBCLT,BCLT}
199: differ from  the physically correct ``no-incoming-radiation'' data
200: is unclear. This crucial issue will be further discussed below.
201: 
202: Recent simulations~\cite{Gpc} based on population synthesis codes predict that
203: $\sim 50 \% -80\%$ of neutron star-black hole (NS-BH) binaries in the Galactic 
204: field may have tilt angles (i.e., the angle between the black hole (BH) spin and the orbital
205: angular momentum) between $0$ and $40^o$ and $10-20 \%$ of
206: NS-BH binaries between $40^o$ and $50^o$. By studying the formation of close compact
207: binaries and the misalignment angle that can occur after the second core-collapse
208: event, Kalogera~\cite{K00} predicted that the majority of BH-BH binaries
209: in Galactic binaries may have a tilt angle
210: smaller than $30^{o}$. All these results assume that the misalignement is entirely due to
211: the recoil velocity (``kick'') imparted to the NS (or the smaller BH in the binary)
212: at birth by the core-collapse.
213: However, the spin properties of NS-BH and BH-BH binaries in dense environment
214: and centers of globular clusters could be very different than in
215: the Galactic field.
216: Considering the low event rates, $\sim 1$ per 2 years, of binary coalescences in
217: first generation of ground-based detectors, it is worthwhile to adopt a conservative
218: point of view and investigate waveforms for generic spin configurations.
219: Little is known about the magnitude of the spin of NSs and BHs.
220: {}From the observed pulsars the dimensionless rotation parameter $a_{\rm NS}$
221: takes values in the range $0.005 - 0.02$. Our analysis will focus on BH-BH binaries and
222: we shall consider arbitrary spins: $0 < a_{\rm BH} < 1$.
223: 
224: For completeness, we investigate the two-body dynamics by adopting two approaches:
225: the straightforward post-Newtonian (PN)-expanded Hamiltonian~\cite{DS88,JS98}
226: and the PN-resummed Hamiltonian \`a la EOB~\cite{BD1}, \cite{DJS}, \cite{TD} [henceforth
227: referred to simply as the EOB-Hamiltonian]. For simplicity, instead of using
228: the Kerr-deformed EOB-Hamiltonian derived by Damour in Ref.~\cite{TD}, we shall use as Hamiltonian
229: in this paper the {\it sum} of the purely orbital (Schwarzschild-deformed) 
230: EOB-Hamiltonian~\cite{BD1,DJS} and of the {\it separate} contributions
231: due to spin-orbit and spin-spin effects. [Note that, among the
232: spin-spin terms, there are the terms due to monopole-quadrupole interactions~\cite{EP}, \cite{TD}.]
233: 
234: The paper is organized as follows. Section~\ref{sec2} summarizes the main 
235: results of the conservative part of the two-body dynamics in the Hamiltonian
236: formalism, and contains formula for the PN-expanded and EOB Hamiltonians
237: up to 3PN order. In Sec.~\ref{sec3} we augment the dynamics with radiation-reaction 
238: effects. We derive the radiation-reaction force which includes spin effects and 
239: matches known rates of energy and angular momentum losses on quasi-circular orbits. 
240: [Our result agrees with the recent results of Will~\cite{Will} that appeared after we had completed our work.]
241: In Sec.~\ref{sec4.1} we define the two-body approximants and discuss
242: the initial conditions used to evolve 
243: the precessing two-body dynamics; we compare the
244: (lowest-order) waveforms
245: obtained using PN-expanded and (PN-resummed) EOB
246: dynamics by computing the overlaps between these two types of
247: waveforms. In view of the greater a priori reliability of the
248: EOB approach, we use them (and only them) to estimate the energy and angular momentum released
249: during the last stages of evolution. Section~\ref{sec5} contains our 
250: main conclusions.
251: 
252: We leave to future work a thorough application of
253: our results to data analysis purposes.
254: 
255: \section{Conservative Hamiltonian including spin-orbit and spin-spin effects}
256: \label{sec2}
257: 
258: \subsection{Orbital third-post-Newtonian expanded Hamiltonian}
259: \label{sec2.1}
260: 
261: The  purely {\it orbital} (non-spinning)third-post-Newtonian Hamiltonian
262: $H^{\rm 0}=H^{\rm 0}(\vX,\vP)$ (in the center of mass frame,
263: and after subtraction of the total rest-mass term)
264: was derived in Ref.~\cite{JS98} (completed by Refs.~\cite{DJSPoincare,DJSd}). In scaled variables, and
265: written as a straigthforward PN-expansion, it reads
266: (see Ref.~\cite{DJS}):
267: %
268: \beq
269: \label{3.1}
270: H^{\rm 0}(\vX,\vP)_{\rm EXP} = \mu\,\widehat{H}^{\rm 0}(\vq,\vp) = \mu\,\left [\widehat{H}_{\rm Newt}(\vq,\vp) +
271: \widehat{H}_{\rm 1PN}(\vq,\vp) +\widehat{H}_{\rm 2PN}(\vq,\vp) +
272: \widehat{H}_{\rm 3PN}(\vq,\vp) \right ]
273: \eeq 
274: %
275: where $\mu = m_1\,m_2/M$, $M=m_1+m_2$ and $(\vq,\vp)$ denote the
276: canonical variables ${\bf p}\equiv {\bf P_1}/\mu=-{\bf P_2}/\mu$, and 
277: ${\bf q} \equiv {\bf X}/M =({\bf X}_1 - {\bf X}_2)/M$, where ${\bf X_1}$ and ${\bf X_2}$ 
278: are the positions of the black hole centers of mass in quasi--Cartesian Arnowitt-Deser-Misner 
279: (ADM) coordinates. The Newtonian term and the PN contributions read
280: (denoting $\eta \equiv \mu/M = m_1m_2/M^2$):
281: %
282: \bea
283: \widehat{H}_{\rm Newt}\left({\bf q},{\bf p}\right) &=& \frac{\pp}{2} -
284: \frac{1}{q}\,, \label{eq:hfirst}\\
285: \widehat{H}_{\rm 1PN}\left({\bf q},{\bf p}\right) &=& \frac{1}{8}(3\eta-1)\ppp^2
286: - \frac{1}{2}\left[(3+\eta)\pp+\eta\np^2\right]\frac{1}{q} + \frac{1}{2q^2}\,,\\
287: \widehat{H}_{\rm 2PN}\left({\bf q},{\bf p}\right)
288: &=& \frac{1}{16}\left(1-5\eta+5\eta^2\right)\ppp^3
289: + \frac{1}{8} \left[
290: \left(5-20\eta-3\eta^2\right)\ppp^2-2\eta^2\np^2\pp-3\eta^2\np^4 \right]\frac{1}{q}
291: \nonumber \\
292: && + \frac{1}{2} \left[(5+8\eta)\pp+3\eta\np^2\right]\frac{1}{q^2}
293: - \frac{1}{4}(1+3\eta)\frac{1}{q^3}\,,
294: \eea
295: %
296: \bea
297: \widehat{H}_{\rm 3PN}\left({\bf q},{\bf p}\right)
298: &=& \frac{1}{128}\left(-5+35\eta-70\eta^2+35\eta^3\right)\ppp^4
299: \nonumber \\
300: && + \frac{1}{16}\left[
301: \left(-7+42\eta-53\eta^2-5\eta^3\right)\ppp^3
302: + (2-3\eta)\eta^2\np^2\ppp^2
303: + 3(1-\eta)\eta^2\np^4\pp - 5\eta^3\np^6
304: \right]\frac{1}{q}
305: \nonumber \\
306: && +\left[ \frac{1}{16}\left(-27+136\eta+109\eta^2\right)\ppp^2
307: + \frac{1}{16}(17+30\eta)\eta\np^2\pp + \frac{1}{12}(5+43\eta)\eta\np^4
308: \right]\frac{1}{q^2} \\
309: && +\left\{ \left[ -\frac{25}{8} + \left(\frac{1}{64}\pi^2-\frac{335}{48}\right)\eta 
310: - \frac{23}{8}\eta^2 \right]\pp
311: + \left(-\frac{85}{16}-\frac{3}{64}\pi^2-\frac{7}{4}\eta\right)\eta\np^2 
312: \right\}\frac{1}{q^3}
313: \nonumber \\
314: && + \left[ \frac{1}{8} + \left(\frac{109}{12}-\frac{21}{32}\pi^2\right)\eta 
315: \right]\frac{1}{q^4}\,, \label{eq:hlast}
316: \eea
317: %
318: where the scalars $q$ and $p$ are the
319: (coordinate) lengths of the two vectors $\vq$ and $\vp$; and the vector ${\bf n}$ is
320: just ${\bf q}/q$.
321: 
322: \subsection{Orbital third-post-Newtonian effective-one-body Hamiltonian}
323: \label{sec2.2}
324: 
325: As was emphasized in previous work (see e.g. \cite{KWW,WS}), and as
326: we shall confirm below, the non-resummed PN-expanded Hamiltonian
327: (or the non-resummed PN-expanded equations of motion) do not lead to
328: a reliable description of the evolution near the last stable
329: circular orbit, nor, {\it a fortiori} during the transition
330: between inspiral and plunge. On the other hand, it was argued
331: in \cite{BD1,BD2,DJS} that the EOB approach
332: defines a specific resummation of the PN-expanded Hamiltonian
333: which leads to a much more reliable description of the dynamical
334: evolution near the last stable circular orbit, and of the
335: transition between inspiral and plunge.
336: 
337: The explicit expression of the purely
338: orbital, EOB-Hamiltonian is~\cite{BD1}:
339: 
340: \beq
341: \label{himpr}
342: H^{\rm 0}_{\rm EOB}(\vX^\prime,\vP^\prime) = M\,\sqrt{1 + 2\eta\,\left ( \frac{H_{\rm eff}(\vX^\prime,\vP^\prime) 
343: -\mu}{\mu}\right )} -M\,.  
344: \eeq
345: %
346: where $H_{\rm eff}$ is given by~\cite{BD1,DJS}:
347: %
348: \bea 
349: \label{eq:genexp}
350: && H_{\rm eff}(\vX^\prime,\vP^\prime) = \mu\, 
351: \widehat{H}_{\rm eff}({\mathbf q}^\prime,{\mathbf p}^\prime) \nonumber \\
352: &=& \mu\,\sqrt{A (q^\prime) \left[ 1 + 
353: {\mathbf p}^{\prime\,2} +
354: \left( \frac{A(q^\prime)}{D(q^\prime)} - 1 \right) ({\mathbf n}^\prime \cdot {\mathbf p}^\prime)^2
355: + \frac{1}{q^{\prime\,2}} \left( z_1 ({\mathbf p}^{\prime\,2})^2 + z_2 \, {\mathbf p}^{\prime\,2}
356: ({\mathbf n}^\prime \cdot {\mathbf p}^{\prime\,})^2 + z_3 ({\mathbf n}^\prime \cdot {\mathbf
357: p}^{\prime})^4 \right) \right]} \,,
358: \eea
359: %
360: with ${\bf q}^\prime$ and ${\bf p}^\prime$ being the reduced canonical variables obtained by rescaling 
361: $\vX^\prime$ and $\vP^\prime$ by $M$ and $\mu$, respectively, ${\bf n}^\prime = \vq^\prime/q^\prime$ 
362: where we set $q^\prime = |{\bf q}^\prime|$. 
363: The coefficients $z_1,z_2$ and $z_3$ are arbitrary, subject to the constraint
364: %
365: \beq 8z_1 + 4z_2 +3z_3 = 6(4-3\eta)\,\eta\,.  
366: \eeq
367: %
368: Setting (as in Ref.~\cite{bcv1})  $z_1 = \eta \tilde{z}_1,
369: z_2 =\eta \tilde{z}_2$ and $z_3 =\eta \tilde{z}_3$, the coefficients $A(q^\prime)$ and $D(q^\prime)$ in Eq.~(\ref{eq:genexp}) read:
370: %
371: \bea
372: \label{coeffA}
373: A(q^\prime) &=& 1 - \frac{2}{q^\prime}+\frac{2\eta}{q^{\prime\,3}}+ \left [ \left (
374: \frac{94}{3}-\frac{41}{32}\pi^2\right ) -\tilde{z}_1\right ]\,\frac{\eta}{q^{\prime\,4}}\,,\\
375: \label{coeffD}
376: D(q^\prime) &=& 1 -\frac{6\eta}{q^{\prime\,2}}+\left [7\tilde{z}_1 +\tilde{z}_2+ (3\eta-26)\right ]\,
377: \frac{\eta}{q^{\prime\,3}}\,. 
378: \eea
379: %
380: As done in Ref.~\cite{DJS}, 
381: we restrict ourselves to the case $\tilde{z}_1=\tilde{z}_2=0$ and improve the behavior
382: \footnote{As shown in \cite{DJS}, the use of the
383: straightforward PN-expanded, 3PN-{\it accurate} ``effective potential''
384: $A(q^\prime)$ does not lead to a well-defined last stable circular orbit 
385: (contrary to what happens in the 2PN-accurate case \cite{BD1}). This
386: is due to the rather large value of the 3PN coefficient 
387: $\frac{94}{3}-\frac{41}{32}\pi^2 \simeq 18.688$ entering the PN expansion
388: of $A(q^\prime)$. Replacing the PN-expanded form of $A(q^\prime)$
389: by a Pad\'e approximant cures this problem.[See also \cite{DGG} for
390: a comparison of the physical consequences of various possible
391:  resummations of $A(q^\prime)$.]} by replacing
392:  the ``effective potential'' $A(q^\prime)$ with the Pad\'e approximants
393: %
394: \beq
395: A_{P_2}(q^\prime) = \frac{q^\prime\,(-4+2q^\prime+\eta)}{2q^{\prime\,2}+2\eta+q^\prime\,\eta}\,,
396: \label{coeffPA2}
397: \eeq
398: %
399: at 2PN order and
400: %
401: \beq
402: \label{coeffPA}
403: A_{P_3}(q^\prime) = \frac{q^{\prime\,2}\,[(a_4(\eta,0)+8\eta-16) + q^\prime\,(8-2\eta)]}{
404: q^{\prime\,3}\,(8-2\eta)+q^{\prime\,2}\,(a_4(\eta,0)+4\eta)+q^\prime\,(2a_4(\eta,0)+8\eta)+4(\eta^2+a_4(\eta,0))}\,,
405: \eeq
406: %
407: at 3PN order where
408: %
409: \beq \label{a4}
410: a_4(\eta,\tilde{z}_1) = \left [\frac{94}{3}-\frac{41}{32}\pi^2
411: -\tilde{z}_1\right ]\,\eta\,.
412: \eeq
413: %
414: 
415: \subsection{Adding spin couplings}
416: \label{sec2.3}
417: 
418: There are several ways of including spin effects in the Hamiltonian
419: dynamics. When considering the PN-expanded form of the
420: orbital Hamiltonian $H^{\rm 0}_{\rm EXP}$,
421: it is natural to add the spin-dependent terms as further additional
422: contributions: $ H_{\rm TOT} = H^{\rm 0}_{\rm EXP} +H^{\rm SPIN}$ .
423: On the other hand, when considering the EOB-resummed
424: form of the Hamiltonian $H^{\rm 0}_{\rm EOB}$,
425: it has been argued in Ref.~\cite{TD} that
426: it is probably better to incorporate most of the spin effects
427: within a suitably generalized (\`a la  Kerr) EOB-Hamiltonian,
428: whose explicit form will be found in \cite{TD}.
429: In the present work, we shall, for technical simplicity,
430: adopt a uniform way of including spin effects. Namely,
431: we shall simply include them a linearly added contributions
432: to the basic (PN-expanded or EOB-resummed) orbital Hamiltonian.
433: We shall verify below that, in the EOB-resummed case, the
434: two different ways (\`a la \cite{TD}, or as in the following
435: equation) of incorporating spin effects lead to
436: very similar physical effects.
437: 
438: Finally, the explicit forms we shall use of the ``spinning''
439: Hamiltonian read:
440: 
441: %
442: \beq
443: H_{\rm EXP}(\vX,\vP,\vS_1,\vS_2) = H^{\rm 0}_{\rm EXP}(\vX,\vP) +
444: H_{\rm SO}(\vX,\vP,\vS_1,\vS_2)+ H_{\rm SS}(\vX,\vP,\vS_1,\vS_2)\,,
445: \label{Hspinexp}
446: \eeq
447: or
448: \beq
449: H_{\rm EOB}(\vX,\vP,\vS_1,\vS_2) =  H^{\rm 0}_{\rm EOB}(\vX,\vP) +
450: H_{\rm SO}(\vX,\vP,\vS_1,\vS_2)+ H_{\rm SS}(\vX,\vP,\vS_1,\vS_2)\,,
451: \label{Hspineob}
452: \eeq
453: 
454: 
455: 
456: where~\cite{BGH,BO,DS88}:
457: %
458: \bea
459: \label{3.2}
460: && H_{SO} = 2\frac{\vS_{\rm eff}\cdot \vL}{R^3}\,, \quad 
461: \vS_{\rm eff} \equiv \left ( 1+ \frac{3}{4}\,\frac{m_2}{m_1} \right )\,\vS_1 + 
462: \left ( 1 + \frac{3}{4}\,\frac{m_1}{m_2} \right )\,\vS_2\,,\\
463: \label{3.3}
464: && H_{SS} = H_{S_1 S_1} + H_{S_1 S_2} + H_{S_2 S_2} =
465: \frac{1}{2R^3}\,\frac{\mu}{M}\,\left [ 3 (\vS_0 \cdot \vN) (\vS_0
466: \cdot \vN) - (\vS_0 \cdot \vS_0) \right ]\,, \\ && \vS_0 = \left ( 1 +
467: \frac{m_2}{m_1} \right )\,\vS_1 + \left ( 1 + \frac{m_1}{m_2} \right
468: )\,\vS_2\,,\\ && H_{S_1 S_2} = \frac{1}{R^3}\,\left [ 3 (\vS_1 \cdot
469: \vN) (\vS_2 \cdot \vN) - (\vS_1 \cdot \vS_2) \right ]\,,\\ && H_{S_1
470: S_1} + H_{S_2 S_2}= \frac{1}{2R^3}\,\left [ 3 (\vS_1 \cdot \vN) (\vS_1
471: \cdot \vN) - (\vS_1 \cdot \vS_1) \right ]\,\frac{m_2}{m_1} +
472: \frac{1}{2R^3}\,\left [ 3 (\vS_2 \cdot \vN) (\vS_2 \cdot \vN) - (\vS_2
473: \cdot \vS_2) \right ]\,\frac{m_1}{m_2} \,, 
474: \eea
475: %
476: with $\vN = \vX/R$, $R = |\vX|$ and $\vL = \vX \times \vP$.
477: The spin-spin term $H_{S_1 S_2}$ was derived in Ref.~\cite{BO}, 
478: while the spin-spin terms $H_{S_1 S_1}, H_{S_2 S_2}$ which are valid {\it only} 
479: for a BH binary were derived in Ref.~\cite{TD} 
480: [see discussion around Eqs.~(2.51)--(2.55) in Ref.~\cite{TD} and also
481: Ref.~\cite{EP}]. 
482: They originate from the interaction of the monopole $m_2$ with the spin-induced 
483: quadrupole moment of the spinning black hole of mass $m_1$ and viceversa. 
484: The spin-induced quadrupole moment of a NS depends
485: on the equation of state. So, if we were applying our approch to 
486: NS binaries we could take into account the 
487: monopole-quadrupole interaction by multiplying  $H_{S_1 S_1}, H_{S_2 S_2}$
488: by some equation-of-state-dependent coefficient $\gamma$ [see Ref.~\cite{EP}].
489: 
490: \subsection{Equations of motion and conserved quantities}
491: 
492: The time evolution of any dynamical quantity 
493: $f(\vX,\vP,\vS_1,\vS_2)$ is given by~\cite{TD}:
494: %
495: \beq
496: \label{3.4}
497: \frac{d }{dt} f(\vX,\vP,\vS_1,\vS_2) = \{f,H\}\,,
498: \eeq
499: %
500: where with $\{...,...\}$ we indicated the Poisson brackets
501: $\{X^i,P_j\}= \delta^i_j$. The Hamilton equations
502: of motion are:
503: %
504: \beq
505: \label{3.5}
506: \frac{d \vX}{d t} = + \frac{\partial H}{\partial \vP}\,, \quad \quad
507: \frac{d \vP}{d t} = - \frac{\partial H}{\partial \vX} \,.
508: \eeq
509: %
510: The equations of motion for the spins are easily derived as~\cite{TD,TH,BO}: 
511: %
512: \bea
513: \label{3.6}
514: \frac{d }{dt} \vS_1 = \{\vS_1,H \} = \frac{\partial H}{\partial \vS_1} \times \vS_1 = 
515: \vO_1 \times \vS_1\,,\\
516: \label{3.7}
517: \frac{d }{dt} \vS_2 = \{\vS_2,H \} = \frac{\partial H}{\partial \vS_2} \times \vS_2 = 
518: \vO_2 \times \vS_2\,,
519: \eea
520: %
521: with
522: %
523: \bea
524: \label{3.8}
525: \vO_1 &=& \left ( 2 + \frac{3}{2}\,\frac{m_2}{m_1} \right
526: )\,\frac{\vL}{R^3} + \frac{1}{R^3}\,\left [3\vN\,(\vS_2 \cdot \vN) -
527: \vS_2 \right ] + \frac{3}{R^3}\,\frac{m_2}{m_1}\,\vN\,(\vS_1 \cdot \vN)
528: \,, \\
529: \label{3.9}
530: \vO_2 &=& \left ( 2 + \frac{3}{2}\,\frac{m_1}{m_2} \right
531: )\,\frac{\vL}{R^3} + \frac{1}{R^3}\,\left [3\vN\,(\vS_1 \cdot \vN) -
532: \vS_1 \right ] + \frac{3}{R^3}\,\frac{m_1}{m_2}\,\vN\,(\vS_2 \cdot \vN)\,.  
533: \eea 
534: %
535: Note that a consequence of the above spin-evolution equations
536: is the constancy of the lengths of the spin vectors:
537: $\vS_1^2 =$ cst., $\vS_2^2 =$ cst.
538: 
539: When using the EOB Hamiltonian Eq.~(\ref{Hspineob})
540:  we should in principle
541: consider the canonical transformation between $(\vX,\vP)$ and 
542: $(\vX^\prime,\vP^\prime)$ which is explicitly given as a PN expansion in 
543: Refs.~\cite{BD1,DJS}.
544: However, since the Hamilton equations are valid in any canonical coordinate 
545: system, when we evolve the EOB dynamics we write the Hamilton equations 
546: in terms of $(\vX^\prime,\vP^\prime)$ and for the spinning part we 
547: neglect the differences between $(\vX^\prime,\vP^\prime)$ and $(\vX,\vP)$ 
548: which are of higher PN order. 
549: When in the following sections we compare the results between PN-expanded and PN-resummed 
550: Hamiltonians, we will always compare quantities
551: which are gauge invariant to lowest order.
552: 
553: One of the advantages of using an Hamiltonian formalism is that
554: one can immediately derive from the fundamental symmetries
555: of the relative dynamics (time translation and spatial
556: rotations) two {\it exact} conservation laws: that of the total
557: energy $E = H$, and that of the total angular momentum
558: $\vJ = \vL + \vS_1 + \vS_2$.
559: Using Eqs.~(\ref{3.6}), (\ref{3.7}) it is easy to
560: check the conservation of the total energy,
561: %
562: \beq
563: \label{3.10}
564: \frac{d H}{d t} = \frac{\partial H}{\partial \vX}\,\frac{d \vX}{d t} +  
565: \frac{\partial H}{\partial \vP}\,\frac{d \vP}{d t} +  \frac{\partial H}{\partial \vS_1}\,
566: \frac{d \vS_1}{d t} + \frac{\partial H}{\partial \vS_2}\,\frac{d \vS_2}{d t}=0\,.
567: \eeq
568: %
569: 
570: Similarly, one easily checks the conservation of the total
571: angular momentum $\vJ = \vL + \vS_1 + \vS_2$. Note the
572: remarkable fact that the orbital contribution to
573:  $\vJ$ is exactly given, at any PN-order, by the ``Newtonian-looking''
574:  expression $\vL = \vX \times \vP$. This is contrary to what
575:  happens when working within a Lagrangian formalism, where the
576:  expression for the conserved orbital angular momentum gets
577:  modified at each PN order: $\vL =\mu \vX \times \vv + O(c^{-2})$.
578:  Here, all the needed PN contributions are included in the
579:  Hamiltonian $H$ (and thereby imply that
580: $ \vV =  \partial H/\partial \vP$ is a complicated function
581: of $\vV$).
582: 
583: \subsection{Spin-orbit interaction and ``spherical orbits''}
584: 
585:  For most of the dynamical evolution, the spin-spin terms are
586:  much smaller than the spin-orbit ones.
587: If we restrict ourselves to spin-orbit interactions, the equation of motion for
588: the orbital angular momentum reads
589: %
590: \beq
591: \label{3.11}
592: \frac{d \mathbf{L}}{d t} = \{\mathbf{L},H\} = 
593: \frac{2}{R^3} \mathbf{S}_{\rm eff} \times \mathbf{L}\,.
594: \eeq
595: %
596: A useful consequence of this approximate evolution law is
597:  that $\vL^2$ is conserved.
598:  Under the same approximation, the total spin $\vS = \vS_1 + \vS_2$ satisfies
599: the equation:
600: %
601: \beq
602: \label{3.12}
603: \frac{d \vS}{d t} = - \frac{2}{R^3}\,\vS_{\rm eff} \times \vL\,.
604: \eeq
605: %
606: The above (approximate) evolution equations exhibit clearly the (exact) conservation
607: of the total angular momentum $\vJ = \vL + \vS$
608: ($d\vJ/dt =0$).
609: 
610: It is also easily checked that
611: $\vS_{\rm eff} \cdot \vL$ is a conserved quantity under the
612: above, approximate evolution equations.
613: Therefore, as emphasized by Damour~\cite{TD}, when
614: only spin-orbit terms are included the orbital dynamics can
615: be reduced to a simple ``radial Hamiltonian''
616:  $H(R,P_R)=H(R,P_R,\vL^2={\rm const.},\vS_{\rm eff} \cdot
617: \vL = {\rm const.})$ describing the radial motion.
618:  Here $P_R \equiv N^i\,P_i$ is the momentum
619: canonically conjugated to $R$ ($\{R,P_R\}=1$).  In this case there
620: exists a class of {\it spherical orbits} satisfying
621: %
622: \beq
623: \label{3.13}
624: R = {\rm const.}\,, \quad \quad P_R = 0\,, \quad \quad 
625: \frac{\partial H(R,P_R=0,\vL^2, \vL \cdot \vS_{\rm eff})}{\partial R}=0\,.
626: \eeq
627: %
628: 
629: \subsection{Characteristics of the last stable spherical orbit (LSSO)}
630: 
631: When spin-spin interactions are included, those
632: spherical orbits no longer exist as exact solutions. However, as
633: spin-spin effects are always smaller than spin-orbit ones, one expects
634: that the above {\it spherical} orbits will play the same important role
635: as the usual {\it circular} orbits in the non-spinning case.
636: In particular, the {\it Last Stable Spherical Orbit} (LSSO) should play
637: the important role of delineating the transition between adiabatic
638: inspiral and plunge.
639: 
640: The LSSO for the spinning conservative dynamics is determined by setting 
641: %
642: \beq
643: \frac{\partial H_0}{\partial R} = 0=\frac{\partial^2 H_0}{\partial R^2}\,,
644: \label{eq:isco}
645: \eeq
646: % 
647: where $H_0(R,P_R,P_\phi, \cdots) = H(R,0,P_\phi, \cdots)$.
648: 
649: The physical characteristics of the LSSO for
650: (aligned or anti-aligned) spinning configurations were studied in detail
651: in Ref.~\cite{TD} within the more fully resummed Kerr-like EOB-Hamiltonian
652: introduced in that reference. It was also shown in \cite{DGG} that the
653: predictions from the latter Kerr-like EOB-Hamiltonian were in good
654: agreement with the numerical results on corotating
655: black hole (BH) binaries obtained by means of the
656: helical Killing vector approach \cite{GGB}. See Table I of \cite{DGG}.
657: [Note that the agreement with EOB is better when one considers the
658: 3PN accuracy.]
659: The latter Table also shows that the numerical results on irrotational
660: BH binaries obtained by means of other approaches based on considering
661: only the initial value problem (e.g. \cite{PTC}) significantly differ
662: both from the numerical helical-Killing-vector (HKV) results, and the
663: analytical EOB-ones. 
664: For recent work improving  the numerical implementation of the HKV approach
665: (which is closely related to the ``conformal thin-sandwich'' decomposition)
666: and confirming that it yields results that agree well with the
667: EOB approach, see~\cite{Cook,CookPfeiffer}. As all the
668: currently published numerical
669: estimates of the physical characteristics of close binaries
670: made of {\it spinning } BH's (such as \cite{PTC}) use
671: initial-value-problem approaches rather than the physically
672: better motivated HKV one, we shall not try to compare here
673: the analytical EOB predictions for spinning configurations
674: with numerical results. On the other hand, it is interesting
675: to compare several different, PN-rooted, analytical approaches
676: in their predictions for the binding energy of close BH binaries.
677: 
678: The most straightforward PN-based analytical approach to the
679: physical characteristics of close BH binaries consists of
680: starting from the fully PN-expanded Hamiltonian (\ref{Hspinexp}),
681: considered as defining an exact dynamics, and then to
682: deduce from it the energy and angular frequency of
683: spherical orbits. [We consider here only configurations
684: where the spins are parallel (or antiparallel) to the orbital
685: angular momentum, so that it makes sense, even in presence of
686: spin-spin interactions, to consider spherical (and actually circular)
687:  orbits.] The binding energy of such ``PN-expanded'' spherical
688:  orbits is plotted in the left panel of Fig.~\ref{fig:eomega} as a function
689:  of the orbital angular frequency, for equal-mass BH binaries.
690:  As we see from this Figure, the straightforward PN-expanded
691:  Hamiltonian does not exhibit any minimum of the
692:  binding energy, i.e. does not lead to any {\it Last} Stable Spherical
693:  Orbit (LSSO) in the non-spinning, or aligned maximally spinning
694:  cases. As the best current numerical results on BH binaries
695:  clearly indicate the existence of such LSSO's, this disqualifies
696:  the use of the fully PN-expanded Hamiltonian (\ref{Hspinexp})
697:  for describing close binaries.
698: 
699: It has, however, been pointed out \cite{LB} that more reasonable results, 
700: close to the numerical HKV results, can be obtained by plotting, instead 
701: of the prediction coming straightforwardly from (\ref{Hspinexp}), 
702: the PN-expansion of the analytically computed function ${E}(\Omega)$ 
703: giving the binding energy $E$ as a function of the orbital frequency $\Omega$.
704: Indeed, one can derive from (\ref{Hspinexp}) the following
705: explicit PN-expansion of the (invariant) function
706: ${E}(\Omega)$ \cite{DJSinvar}, \cite{DJS}, \cite{LB}:
707: %
708: \bea
709: \label{s1}
710: E_{\rm 2PN}(\Omega) &=& -\frac{\mu}{2}\,(M\Omega)^{2/3}\,\left \{
711: 1 - \frac{(9+\eta)}{12}\,(M\Omega)^{2/3} + \frac{8}{3} \frac{\hL\cdot \vS_{\rm eff}}{M^2}\,
712: (M\Omega)+
713: \frac{1}{24}(-81 + 57 \eta - \eta^2)\,(M\Omega)^{4/3} \right . \nonumber \\
714: && \left. + \frac{1}{\eta}\,\frac{1}{M^4}\,
715: \left [\vS_1\cdot \vS_2 - 3 (\hL \cdot \vS_1) (\hL \cdot \vS_2)
716:  \right ]\,(M\Omega)^{4/3} \right \}\,,\\
717: \label{s2}
718: E_{\rm 3PN}(\Omega) &=& E_{\rm 2PN}(\Omega) -\frac{\mu}{2}\,(M\Omega)^{2/3}\,\left \{
719: \left [-\frac{675}{64} + \left (\frac{34445}{576}-\frac{205}{96}\pi^2 \right )\eta
720: -\frac{155}{96}\eta^2-\frac{35}{5184}\eta^3 \right ]\,(M\Omega)^{2}\right \}\,.
721: \eea
722: %
723: These functions are plotted in the right panel of Fig.~\ref{fig:eomega}. 
724: Visibly, they are much better behaved than the results plotted on the left 
725: panel, which came directly from the PN-expanded Hamiltonian. 
726: They are also close to the numerical HKV results~\cite{LB}. The fact 
727: that two expressions, which can both be called ``PN-expanded'', and which 
728: are supposedly equivalent ``modulo higher PN terms'', lead to physically 
729: markedly different results lead to conclude that the PN-expanded Hamiltonian 
730: cannot be used to describe the transition from adiabatic inspiral to plunge. 
731: Let us also recall that if we consider, in absence of spins,  
732: the test-mass limit $\eta \to 0$, instead of the equal-mass one $\eta \to 1/4$,  the
733: PN-expansions (\ref{s1}),(\ref{s2}) have been shown (see
734: \cite{DJS}) to be quite inaccurate representations of the known exact expression
735: of the function ${E}(\Omega)$. Indeed, the 2PN-accurate function
736: (\ref{s1}) predicts in this limit an LSO frequency which is
737: $82\%$ larger than the exact one, while the 3PN-accurate one
738: (\ref{s2}) predicts an LSO frequency $27\%$ larger than the exact one.
739: 
740: By contrast with these problematic features of PN-expanded
741: results \footnote{Let us recall here that, in order to be able to describe
742: the transition between inspiral and plunge, we cannot use just
743: the function ${E}(\Omega)$, but we need a full description of the
744: binary dynamics. Therefore, if we wanted to confine ourselves
745: to a ``PN-expanded'' approach, we would have to use either
746: the PN-expanded Hamiltonian (\ref{Hspinexp}), or the corresponding
747: (appropriately truncated) PN-expanded equations of motion.
748: The left panel of Figure \ref{fig:eomega} shows that this would not be
749: a reliable thing to do. This is also clear from some of the
750: results of Ref.\cite{bcv1}.}, the EOB-approach leads to
751: uniformly better behaved results (even if we use it not
752: in the Kerr-like form advocated in \cite{TD}, but in the
753: form (\ref{Hspineob}) used in the present paper). We show
754: in Fig.~\ref{fig:eomegaEOB} the EOB analog of Fig.~\ref{fig:eomega}, 
755: i.e. the function ${E}(\Omega)$ deduced from the EOB Hamiltonian 
756: (\ref{Hspineob}) in the (anti-)aligned case. In this EOB case, we  have none of
757: the problems entailed by the ``PN-expanded'' approach, and the
758: uniquely defined curve ${E}(\Omega)$ was shown in \cite{DGG}
759: to agree well with the HKV  numerically determined curve
760: (for corotating holes). Note, however, that {\it aligned}
761: maximally rotating holes lead to a curve which, especially
762: in the 3PN case, reaches a minimum (not shown on Figure \ref{fig:eomegaEOB})
763: only for a rather high angular velocity.
764: 
765: This property of the aligned configurations (as well as
766: the significant difference between the 2PN-EOB result
767: and the 3PN-EOB one) was already emphasized in \cite{TD}.
768: As it will be important for the present paper, 
769: we study it further by plotting in Fig.~\ref{fig:e-a} 
770: the dependence on the $\vL$- projected effective spin parameter
771: %
772: \beq
773: \chi_L \equiv \frac{\vS_{\rm eff} \cdot \hL}{M^2}
774: \eeq
775: %
776: (where $\vS_{\rm eff}$ was defined in Eq.~(\ref{3.2}) above) of the binding
777: energy $E$ and the angular frequency $\Omega$ {\it at the LSSO},
778: i.e. at the minimum of the ${E}(\Omega)$ curve.
779: This Figure shows four results obtained for
780: equal-mass and equal-spin configurations within
781: the EOB approach: (i) the result obtained from the
782: Hamiltonian (\ref{Hspineob}) when using the 2PN-accurate
783: orbital EOB Hamiltonian, (ii) the result obtained from
784: (\ref{Hspineob}) when using the 3PN-accurate
785: orbital EOB Hamiltonian, (iii) the result obtained from
786: the  2PN-accurate Kerr-like Hamiltonian introduced in \cite{TD},
787: and (iv) the result obtained from the  3PN-accurate
788: Kerr-like Hamiltonian introduced in \cite{TD}. [The latter two
789: Hamiltonians are referred to in the caption as ``SO Resummed'',
790: because they include a resummation \`a la EOB of the spin-orbit
791: interactions.] In addition, as we cannot show
792: on this plot the minimum of the ${E}(\Omega)$ curve deduced
793: from the PN-expanded Hamiltonian, because the left panel
794: of Fig.~\ref{fig:e-a} shows that it does not exist, we show instead,
795:  for comparison purposes, the characteristics
796: of the  minimum of the PN-expanded functions (\ref{s1}), (\ref{s2})
797: [i.e. the right panel of Fig.~(\ref{fig:eomega})].
798: 
799: It is interesting to note on Fig.~\ref{fig:e-a} that the effect of
800: resumming (\`a la Kerr) or not the spin-orbit interaction
801: seems to be rather small. We see also that, when
802: considering anti-aligned configurations, all calculations
803: give very similar results. This is not surprising as
804: the {\it attractive}  ($H_{SO} <0$) nature of the {\it anti-aligned}
805: spin-orbit (and spin-spin) interaction has the effect
806: of pushing the LSSO {\it outwards}, i.e. toward larger-radius,
807: lower-frequency, less bound and therefore less relativistic
808:  configurations. On the other hand, working
809: at the 2PN or the 3PN level induces,as already pointed out in \cite{TD},
810:  a very significant difference for the LSSO characteristics in the {\it aligned} case (positive
811: $\chi_L$). In this case, because of the {\it repulsive}($H_{SO} >0$)
812: nature of the {\it aligned} spin-orbit (and spin-spin) interaction, the LSSO is drawn
813: towards closer, higher-frequency, more bound and more
814: relativistic configurations. For such very compact
815: configurations the {\it repulsive} sign ($a_4 >0$)
816: of the 3PN contribution to the effective potential $A(q)$
817: further amplifies, by a ``snow-ball effect'', the tendency
818: toward closer, and more bound configurations.
819: We think that this could be a physically real effect due (as confirmed independently by Refs.~\cite{BDE,Itoh}) 
820: to the large positive value of the crucial 3PN coefficient entering  $a_4$, Eq.~(\ref{a4}).  
821: This large positive value for $a_4$ is also needed to improve (with respect to the 2PN case)
822: the  agreement between the HKV corotating results and the 3PN EOB ones \cite{DGG}.
823: It would be interesting to have numerical HKV studies of 
824: the LSSO for moderately- and fast-spinning aligned BH's to test the predictions made by the EOB approach.
825: [The less reliable numerical results of the initial-value-problem of Ref.\cite{PTC},
826: which extend up to mildly positive values of $\chi_L \sim 0.17$
827: are in rough qualitative agreement (especially 
828: for the dependence of $\Omega_{LSSO}$ on $\chi_L$) with the
829: EOB predictions, but their quantitative agreement is too
830: poor to reach a firm conclusion).]
831: 
832: Let us note in passing that the significant dependence of the
833: LSSO frequency on the effective spin parameter $\chi_L$
834: makes it desirable for data-analysis purposes, when one is content
835: with using adiabatic templates~\cite{bcv2}, to use at least
836: templates whose ending frequency is not fixed say to the usually considered 
837: Schwarzschild LSO, but varies with masses and spins as suggested by the EOB 
838: approach.
839: 
840: Finally, another consequence of the significant dependence of the
841: LSSO frequency on the effective spin parameter $\chi_L$, is 
842: drawn in Fig.~\ref{snr}, where we compare the signal-to-noise ratios (SNRs) as 
843: function of the binary total mass for an optimally oriented equal-mass binary at 100 Mpc. 
844: We use LIGO design sensitivity noise curve~\cite{DIS98}.  
845: The SNRs are obtained  observing the inspiral from 40 Hz until the LSSO predicted by 
846: the EOB approach at 3PN order. The three curves refer to 
847: non-spinning binaries and binaries with $\chi_L = -0.875, 0.25$. 
848: Figure~\ref{snr} reveals a bias towards 
849: first detecting mostly aligned spinning binaries with high masses, 
850: as pointed out in \cite{TD}.
851: 
852: \section{Radiation reaction, including spin-effects}
853: \label{sec3}
854: 
855: The previous section has reviewed various ways of describing the
856: conservative dynamics of binary systems (including spin effects). In
857: the present section, we discuss the inclusion of radiation reaction
858: effects, with emphasis on determining the spin-modifications of
859: radiation reaction. Within the Hamiltonian approach, that we use here,
860: radiation reaction can be incorporated by modifying the usual Hamilton
861: equations in the following way
862: $$
863: \frac{dX^i}{dt} = \{ X^i , H \} = \frac{\partial H}{\partial P_i} \, ,
864: $$
865: %
866: \beq
867: \label{n1}
868: \frac{dP_i}{dt} = \{ P_i , H \} + F_i = - \frac{\partial H}{\partial X^i} + F_i \, .
869: \eeq
870: %
871: Here, $F_i$ denotes a ``non-conservative force'', which is added to
872: the evolution equation of the (relative) momentum to take into account
873: radiation-reaction (RR) effects. This Radiation-Reaction (RR)
874:  force $\vF$ depends, a
875: priori, both on the (relative) orbital variables $\vX$, $\vP$ and on
876: the spin variables $\vS_1$, $\vS_2$. In the present paper, our aim
877: will be limited to determining $F_i$ under the following two
878: simplifying assumptions: (i) we consider only {\it quasi-circular
879: orbits}, and (ii) we shall retain only the {\it leading spin-dependent
880: terms}. After the completion of the work reported in this section,
881: there appeared a work of Will \cite{Will} dealing with
882: spin-dependent radiation reaction effects in general orbits.
883: As the derivations are not the same, and yield results which
884: we have checked to be equivalent (for circular orbits), but
885: expressed in different variables
886: (Hamiltonian $\vX$, $\vP$ here vs. Lagrangian $\vX$, $\vV$
887: for \cite{Will}), we feel it worth to briefly report our
888: derivation.
889: 
890: Because of the assumption (ii) we look for terms in $F_i$
891: which are {\it linear} in the spin-tensors $S_{ij}^a \equiv
892: \varepsilon_{ijk} \, S_a^k$ ($a=1,2$) of the two considered compact
893: bodies. [Note that spin effects enter the metric only through the spin
894: tensors $S_{ij}^a$, rather than through the axial spin vectors
895: $S_a^k$.] As the time derivative of $S_{ij}^a$ contains a ``small''
896: post-Newtonian factor $G/c^2$, the leading spin-dependent terms in
897: $F_i$ will contain only the undifferentiated spin tensors.
898: 
899: 
900: Using Euclidean invariance, the spin-dependent terms in the force $F_i
901: (\vX , \vP , \vS_a)$ must be a combination of three types of
902: contributions: $c_1 \, S_{ij} \, X^j$, $c_2 \, S_{ij} \, P^j$ and $c_i
903: \, S_{jk} \, X^j P^k$, where $c_1 (\vX , \vP)$, $c_2 (\vX , \vP)$ are
904: some scalar functions of $\vX$, $\vP$, while $c_i$ is a vector
905: function of $\vX$, $\vP$. [Here, $S_{ij}$ denotes one of the two spin
906: vectors. We shall sum over the two possible spins at the end.]
907: Imposing that the radiation reaction force $F_i$ be {\it odd under
908: time reversal}, i.e. odd under the simultaneous changes $X^i \to X^i$,
909: $P_i \to -P_i$, $S_{ij} \to - S_{ij}$, tells us that: $c_1 (\vX ,
910: \vP)$ must be an {\it even} function of $\vP$, $c_2 (\vX , \vP)$ must
911: be an {\it odd} function of $\vP$, and the vector $c_i (\vX , \vP)$
912: must be an {\it odd} function of $\vP$. [Note also that $c_i$ must be
913: a true vector, not an axial vector. By parity invariance no
914: $\varepsilon_{ijk}$ can appear, except in combination with $S_a^k$.]
915: Therefore, if we further decompose $c_i = c_3 (\vX , \vP) \, P_i + c_4
916: (\vX , \vP) \, X_i$, the coefficient $c_3 (\vX , \vP)$ must be {\it
917: even} in $\vP$, while the coefficient $c_4 (\vX , \vP)$ must be {\it
918: odd} in $\vP$.
919: 
920: At this point, our simplifying assumption (i) above (quasi-circular
921: motion) will bring a drastic simplification. Indeed, a time-odd scalar
922: must contain an odd power of the combination $X^k P_k$. However, this
923: combination vanishes along circular orbits (and is therefore
924: subleadingly small along adiabatically inspiralling orbits). To
925: leading order the scalar coefficients $c_2$ and $c_4$ therefore
926: vanish, and we conclude that $F_i$ contains only two independent spin
927: contributions: $c_1 (\vX , \vP) \, S_{ij} \, X^j + c_3 (\vX , \vP) \,
928: P_i \, S_{jk} \, X^j P^k$. It will be convenient in the following to
929: further decompose the vector $S_{ij} \, X^j$ entering the first
930: contribution (which is orthogonal to $X^i$) into its component along
931: the direction of $P_i$, and its component orthogonal to $P_i$, say
932: %
933: \beq
934: \label{n2}
935: (S_{ij} \, X^j)^{\perp} \equiv (\delta_{ik} - P_i \, P_k / \vP^2) \,
936: S_{kj} \, X^j = S_{ij} \, X^j + \frac{P_i}{\vP^2} \, S_{jk} \, X^j P^k
937: \, .  \eeq
938: %
939: It is easily checked that, along circular orbits $(X^i \, P_i = 0)$,
940: the vector (\ref{n2}) is orthogonal {\it both} to $\vP$ and to
941: $\vX$. Therefore, $(S_{ij} \, X^j)^{\perp}$ is parallel to the orbital
942: angular momentum (axial) vector
943: %
944: \beq
945: \label{n3}
946: L_i \equiv \varepsilon_{ijk} \, X^j \, P_k \, .
947: \eeq
948: %
949: It is easily checked that
950: %
951: \beq
952: \label{n4}
953: (S_{ij} \, X^j)^{\perp} = \frac{R^2}{\vL^2} \, (\vP \cdot \vS) \, L_i
954: = \frac{1}{\vP^2} \, (\vP \cdot \vS) \, L_i \, .  \eeq
955: %
956: 
957: Finally, adding the usual spin-independent radiation reaction
958: (parallel to $P_i$ for circular orbits), and summing over the two
959: bodies, we conclude that the RR force can be written as
960: %
961: \beq
962: \label{n5}
963: F_i (\vX , \vP , \vS_a) = B \, P_i + \sum_{a = 1,2} A_a (S_{ij}^a \,
964: X^j)^{\perp} = B \, P_i + \sum_{a=1,2} \frac{A_a}{\vP^2} \, (\vP \cdot
965: \vS_a) \, L_i \, , \eeq
966: %
967: with
968: %
969: \beq
970: \label{n6}
971: B \equiv B_0 + \sum_{a=1,2} C_a \, S_{jk}^a \, X^j P^k = B_0 + \sum_{a=1,2} C_a \, \vL \cdot \vS_a \, ,
972: \eeq
973: %
974: where $B_0$, $C_a$ and $A_a$ are some time-even functions of $\vX$ and $\vP$.
975: 
976: To determine the coefficients $B_0$, $C_a$ and $A_a$, we now impose
977: that there be a balance between the losses of mechanical energy and
978: angular momentum of the system due to the additional force $F_i$ in
979: the Hamilton equations of motion (\ref{n1}) and the losses of energy
980: and angular momentum at infinity due to the emission of gravitational
981: radiation. Let us first recall that, in the Hamiltonian formalism, the
982: quantities
983: $$
984: E(t) \equiv H(\vX (t) , \vP (t) , \vS_a (t)) \, ,
985: $$
986: %
987: \beq
988: \label{n7}
989: J_{ij} (t) \equiv X^i \, P_j - X^j \, P_i + S_{ij}^1 + S_{ij}^2 \, ,
990: \eeq
991: %
992: are exact constants of the motion in absence of RR force in
993: Eqs. (\ref{n1}) and in the Hamiltonian equations for spin
994: evolution.
995: When adding the RR force $\vF$ in Eqs. (\ref{n1}) (and no
996: corresponding RR torque in the spin evolution equations) we find that
997: $E$ and $\vJ$ evolve as
998: %
999: \beq
1000: \label{n8a}
1001: \frac{dE}{dt} = \frac{\partial H}{\partial P_i} \, F_i = \dot X^i \, F_i \, ,
1002: \eeq
1003: %
1004: %
1005: \beq
1006: \label{n8b}
1007: \frac{d \, J_{ij}}{dt} = X^i \, F_j - X^j \, F_i \, .
1008: \eeq
1009: %
1010: Inserting in Eqs. (\ref{n8a},\ref{n8b}) the expression (\ref{n5}) for
1011: the RR force $F_i$, we can easily evaluate the {\it averaged} losses
1012: of energy and angular momentum. Along (quasi) circular orbits the
1013: various scalar coefficients $B_0$, $C_a$, $A_a$ are time-independent
1014: (because all basic scalars, $\vX^2 , \vP^2 , \vX \cdot \vP = 0$, are
1015: constant). One then finds that $dE/dt$ is time-independent, while $d
1016: \, J_{ij} / dt$ depends on the orbital phase only in spin-dependent
1017: terms and through the tensor $X^i \, P_j$. Decomposing the latter
1018: tensor into
1019: %
1020: \beq
1021: \label{n9}
1022: X^i \, P_j = \frac{1}{2} (X^i \, P_j + X^j \, P_i) + \frac{1}{2} (X^i
1023: \, P_j - X^j \, P_i) \simeq \frac{d}{dt} \left( \frac{1}{2} \, \mu \,
1024: X^i \, X^j \right) + \frac{1}{2} \, L_{ij} \, , \eeq
1025: %
1026: one easily sees that its orbital average is simply $\langle X^i \, P_j
1027: \rangle = \frac{1}{2} \, L_{ij}$. [We consider averages over the
1028: orbital period, considering all more slowly evolving quantities, such
1029: as $\vL$, as fixed during one orbital period.]
1030: 
1031: When evaluating Eq. (\ref{n8a}) along circular orbits, we cannot use
1032: the Newtonian approximation $\dot X^i \simeq P^i / \mu$ because we
1033: wish to obtain the coefficient $B$ with a high post-Newtonian
1034: accuracy. However we can instead use $\dot X^i \, P_i = \dot\phi \,
1035: P_{\phi} = \omega \, \vert \vL \vert$ where $\omega = \dot\phi =
1036: V/R$ denotes the orbital angular frequency. We finally obtain
1037: %
1038: \beq
1039: \label{n10a}
1040: \frac{dE}{dt} = B \, \omega \, \vert \vL \vert \, ,
1041: \eeq
1042: %
1043: 
1044: %
1045: \beq
1046: \label{n10b}
1047: \left\langle \frac{d \vJ}{dt} \right\rangle = B \, \vL - \frac{1}{2}
1048: \sum_{a=1,2} A_a \, R^2 \, [\vS_a - \vlb (\vlb \, \vS_a)] \, , \eeq
1049: %
1050: where $\vlb \equiv \vL / \vert \vL \vert$ denotes the unit vector
1051: along the orbital angular momentum. Note that Eqs. (\ref{n10a}),
1052: (\ref{n10b}) predict a link between energy loss and angular momentum
1053: loss, namely
1054: %
1055: \beq
1056: \label{n11}
1057: \frac{dE}{dt} = \omega \, \vlb \cdot \left\langle \frac{d \vJ}{dt} \right\rangle \, .
1058: \eeq
1059: %
1060: 
1061: To obtain the values of the coefficients $B$ and $A_a$ (and to test
1062: the prediction (\ref{n11})), we need to compare Eqs. (\ref{n10a}),
1063: (\ref{n10b}) with the values of the averaged fluxes of energy and
1064: angular momentum at infinity. The spin contributions to the latter
1065: losses have been computed by Kidder~\cite{K}. However, one must be
1066: careful with the fact that Kidder expressed most of his results in
1067: terms of {\it harmonic coordinates}, with the choice of a covariant
1068: spin supplementary condition: $S_{\mu\nu}^a \, u_a^{\nu} = 0$.
1069: 
1070: First, using the results of Ref.~\cite{K} as they are, one
1071: straightforwardly checks that the relation (\ref{n11}) is
1072: satisfied. This is a check that it is enough to include RR effects in
1073: the orbital equations of motion (\ref{n1}), without modifying the spin
1074: equations of motion. [For a direct dynamical check, see \cite{Will}.]
1075:  Then, to obtain the value of the coefficient $B$
1076: we can simply use the result (\ref{n10a}), namely
1077: %
1078: \beq
1079: \label{n12}
1080: B = \frac{1}{\omega \, \vert \vL \vert} \ \frac{dE}{dt} \, ,
1081: \eeq
1082: %
1083: where it remains to express $dE/dt$ (computed as a flux at infinity,
1084: using Ref.~\cite{K}) in terms of our basic (Hamiltonian) dynamical
1085: variables. One way to proceed would be to transform the
1086: harmonic-coordinates results of \cite{K} into ADM coordinates (with
1087: the corresponding spin condition $S_{i0} + \frac{1}{2} \, S_{ij} \,
1088: v^j = 0$ \cite{DS88}). The transformation linking the two coordinates
1089: has been worked out in \cite{DS88} (for the spin-dependent terms) and
1090: in \cite{DJS01,ABF01} for the spin-independent parts. However, a
1091: simpler way to proceed is to eliminate references to specific
1092: coordinates by expressing $dE/dt$ (for circular orbits) in terms of
1093: the gauge-invariant orbital frequency $\omega$. Adding also, for
1094: better accuracy the recently completed 3PN flux contribution
1095: \cite{BIJ02,B98,BDEI04}, we have
1096: %
1097: \beq
1098: \label{n13}
1099: \frac{dE}{dt} = - \frac{32}{5} \, \eta^2 \, v_{\omega}^{10} \{ 1 + f_2
1100: (\eta) \, v_{\omega}^2 + [f_3 (\eta) + f_{3{\rm SO}}] \, v_{\omega}^3
1101: + [f_4 (\eta) + f_{4{\rm SS}}] \, v_{\omega}^4 + f_5 (\eta) \,
1102: v_{\omega}^5 + f_6 (\eta) \, v_{\omega}^6 + f_{\ell 6} \, v_{\omega}^6
1103: \, \ln (4 v_{\omega}) + f_7 (\eta) \, v_{\omega}^7 \} \, , \eeq
1104: %
1105: where $v_{\omega} \equiv (GM \omega)^{1/3}$, where the
1106: spin-independent flux coefficients $f_2 (\eta) , \ldots , f_7 (\eta)$,
1107: are given by
1108: %
1109: \beq
1110: \label{n14a}
1111: f_2 (\eta) = - \frac{1247}{336} - \frac{35}{12} \, \eta \, ,
1112: \eeq
1113: %
1114: %
1115: \beq
1116: \label{n14b}
1117: f_3 (\eta) = 4\pi \, ,
1118: \eeq
1119: %
1120: \beq
1121: f_4(\eta) = -\frac{44711}{9072}+\frac{9271}{504}\eta+\frac{65}{18}\eta^2\,, \quad \quad 
1122: f_5(\eta) = -\left ( \frac{8191}{672} + \frac{583}{24}\eta \right )\pi\,, 
1123: \eeq
1124: %
1125: \beq
1126: f_6(\eta) = \frac{6643739519}{69854400}
1127: +\frac{16}{3}\,\pi^2 -\frac{1712}{105}\,\gamma_E  +
1128: \left (-\frac{134543}{7776}+\frac{41}{48}\,\pi^2\right )\eta 
1129: -\frac{94403}{3024}\,\eta^2 - \frac{775}{324}\,\eta^3 \,,
1130: \eeq
1131: %
1132: \beq
1133: f_{\ell 6} = -\frac{1712}{105} 
1134: \eeq
1135: %
1136: \beq
1137: f_7(\eta) = \left (-\frac{16285}{504}+\frac{214745}{1728}\eta+\frac{193385}{3024}\eta^2 \right )
1138: \,\pi\,,
1139: \eeq
1140: %
1141: with $\gamma_E $ being Euler's gamma, and where the spin-dependent
1142: corrections to the latter flux coefficients are
1143: %
1144: \beq
1145: \label{n15a}
1146: f_{3{\rm SO}} = - \left( \frac{11}{4} + \frac{5}{4} \, \frac{m_2}{m_1}
1147: \right) \frac{\vlb \cdot \vS_1}{GM^2} - \left( \frac{11}{4} +
1148: \frac{5}{4} \, \frac{m_1}{m_2} \right) \frac{\vlb \cdot \vS_2}{GM^2}
1149: \, , \eeq
1150: %
1151: %
1152: \beq
1153: \label{n15b}
1154: f_{4{\rm SS}} = \frac{\eta}{48 G^2 m_1^2\,m_2^2} \,
1155: [ 289 (\vlb \cdot \vS_1) (\vlb \cdot \vS_2) - 103 \, \vS_1 \cdot
1156: \vS_2 ] + \mathcal{O}(\vS_1^2) + \mathcal{O}(\vS_2^2)\,.
1157: \eeq
1158: %
1159: The present work was aimed at determining the leading spin-dependent
1160: terms, i.e. the ones linear in $\vS_1$ and $\vS_2$, as exemplified in
1161: the correction $f_{3{\rm SO}}$, Eq. (\ref{n15a}), to the coefficient
1162: $f_3 = 4\pi$. For completeness, as the link (\ref{n12}) between the
1163: ``longitudinal'' part of RR, $F_i^{\rm long} = B \, P_i$, and the
1164: energy loss, is clearly general, we have also used Kidder's results
1165: \cite{K} to write down the part of $B$ which depends on the product
1166: $\vS_1^i \, \vS_2^j$. The numerically similar
1167: contributions which depend on $\vS_1^i \, \vS_1^j$ and $\vS_2^i \,
1168: \vS_2^j$ have not yet been determined. Only partial results
1169: are known. For instance, Poisson \cite{EP} has derived a
1170: contribution to $f_{4{\rm SS}}$ of the form
1171: 
1172: \beq
1173: \left[\frac{3(\vlb \cdot \vS_1)^2 -\vS_1^2}{G^2 m_1^2\,M^2}
1174: +\frac{3 (\vlb \cdot \vS_2)^2-\vS_2^2}{G^2 m_2^2 M^2}
1175: \right]
1176: \,.
1177: \eeq
1178: but many other additional contributions $\mathcal{O}(\vS_1^2) + \mathcal{O}(\vS_2^2)$ have not yet been computed.
1179: 
1180: \medskip
1181: 
1182: Let us finally turn to the determination of the other spin-related
1183: coefficients, $A_a$, in Eq. (\ref{n10b}). Again, we have the technical
1184: problem that Ref.~\cite{K} expressed its results in terms of
1185: harmonic-coordinate quantities. Namely, Eq. (4.11) of Ref.~\cite{K}
1186: expresses the total angular momentum loss $d \vJ / dt$ in terms of the
1187: harmonic distance $r$ and of the harmonic-coordinate ``Newtonian
1188: orbital momentum'' $\vL_N \equiv \mu \, \vx \times \vv$ (where $\vx$
1189: and $\vv$ denote the relative harmonic position and velocity). A
1190: simple way to convert this result to our ADM distance $R$ and our ADM
1191: total orbital momentum $\vL \equiv \vX \times \vP$ is to relate
1192: $\vL_N$ to $\vL$ by comparing the expression (4.7) of Ref.~\cite{K}
1193: for the (gauge-invariant) conserved total angular momentum $\vJ$ with
1194: the corresponding simple ADM expression (\ref{n7}). This yields a
1195: result of the form
1196: %
1197: \beq
1198: \label{n16}
1199: \hat\vL_N \equiv \frac{\vL_N}{\vert \vL_N \vert} = c \, \vL + \left(
1200: \frac{GM}{r} \right)^{\frac{3}{2}} \sum_{a=1,2} \chi_a \, \hat\vs_a
1201: \left( \frac{m_a^2}{M^2} + \frac{1}{4} \, \eta \right) \, , \eeq
1202: %
1203: where the coefficient $c$ is not needed for our present purpose, and
1204: where, following the notation of \cite{K}, $\vS_a \equiv \chi_a \,
1205: m_a^2 \, \hat\vs_a$. Inserting Eq. (\ref{n16}) in Eq. (4.11) of
1206: \cite{K} allows one to compute easily the part of $d \vJ / dt$ which
1207: is proportional to the projection of $\vS_a$ orthogonally to $\vL$.
1208: 
1209: This yields the following expression for the coefficients $A_a$ in Eq. (\ref{n5})
1210: %
1211: \beq
1212: \label{n17}
1213: A_a = \frac{8}{15} \, \eta^2 \, \frac{v_{\omega}^8}{R^3} \left( 61 + 48 \, \frac{m_{a'}}{m_a} \right) \, ,
1214: \eeq
1215: %
1216: where $a' \ne a$ (e.g. $a' = 2$ when $a=1$). Summarizing, the
1217: radiation reaction force to be added to the Hamiltonian equations of
1218: motion (\ref{n1}) reads
1219: %
1220: \beq
1221: \label{n18}
1222: F_i = \frac{1}{\omega \, \vert \vL \vert} \, \frac{dE}{dt} \, P_i +
1223: \frac{8}{15} \, \eta^2 \, \frac{v_{\omega}^8}{\vL^2 R} \left\{
1224: \left(61 + 48 \, \frac{m_2}{m_1} \right) \vP \cdot \vS_1 + \left( 61 +
1225: 48 \, \frac{m_1}{m_2} \right) \vP \cdot \vS_2 \right\} L_i \, , \eeq
1226: %
1227: 
1228: where the energy loss (expressed in terms of the orbital frequency
1229: $\omega$, or equivalently of $v_{\omega} \equiv (GM \omega)^{1/3}$,
1230: and of the spin variables) is given by Eqs. (\ref{n13})--(\ref{n15b}).
1231: We have checked that, after taking into account the relation between
1232: the Hamiltonian variables $\vX, \vP$ and the Lagrangian ones $\vX, \vV$
1233: (which involves spin-dependent terms because of the
1234: first Eq.~(\ref{3.5}), Eq.~(\ref{n18}) agrees with the circular
1235: limit of Eq.~(1.6) of \cite{Will} (which assumes the same spin
1236: condition as we do).
1237: 
1238: Refs.~\cite{DIS98,BD2,DIS01,PS} have shown that (at least 
1239: in the test-mass limit where one can compare analytical and numerical estimates)
1240: it is generally advantageous to replace the Taylor series
1241: in curly brackets on the right hand side (R.H.S.) of Eq. (\ref{n13})
1242: by its (suitably defined) Pad\'e resummation.
1243: In particular, Porter and Sathyaprakash \cite{PS} have compared
1244: ``Taylor'' and ``Pad\'e''  approximants for the flux function 
1245: of a test particle around a Kerr black hole with the exact numerical 
1246: estimates~\cite{fluxS} and concluded
1247: that Pad\'e approximants are, when considering
1248: all values of the spin parameter, both more {\it effectual}
1249: (i.e., larger overlaps with the exact signal) and more
1250: {\it faithful} (i.e., smaller biases in parameter estimates)
1251: than Taylor approximants. [We use here the terminology
1252: introduced in \cite{DIS98}.] In view of this,
1253: and as was already advocated in \cite{BD2}, we consider that
1254: that the best way to incorporate a radiation reaction force in
1255: the EOB approach is to
1256: insert Pad\'e approximants of the flux function (R.H.S.
1257: of (\ref{n13})) in (\ref{n18}). However, for added generality,
1258: we shall also consider the case where we leave the flux function
1259: as a plain Taylor series. Note that, when considering arbitrary
1260: values of the dimensionless spin parameters for the two holes
1261: $\chi_1 \equiv S_1 / G m_1^2$, $ \chi_2 \equiv S_2 / G m_2^2$
1262: we used the normal ``direct'' (i.e., lower-diagonal) Pad\'e-approximants, 
1263: instead of the ``inverse''(i.e., upper-diagonal) ones used in \cite{PS}. 
1264: For some values of the spin parameters both the lower and upper diagonal 
1265: Pad\'e-approximants have poles. When this occurs, we apply the Pad\'e-approximant 
1266: only to the non-spinning part of the flux and add the spinning terms separately. 
1267: There exist other Pad\'e-approximants in which poles are absent and it 
1268: would be very desirable to determine them in the entire parameter space. 
1269: This is beyond the scope of this paper. 
1270: 
1271: In Figs.~\ref{Fig6}, \ref{Fig7} we show T- and (lower-diagonal) P-approximants at 3.5PN 
1272: order for an equal-mass binary and several values of the dimensionless spin 
1273: parameter $\chi=\chi_1=\chi_2$. We notice that the T- and P-approximants are much 
1274: closer in the anti-aligned cases than in the aligned one. 
1275: Since the calculation of the non-spinning flux 
1276: function at 3.5PN order has been completed only recently~\cite{BDE}, 
1277: in Fig.~\ref{Fig5} we contrast the T- and (lower-diagonal) P-approximants at 3PN and 
1278: 3.5PN order for an equal-mass non-spinning binary.
1279: 
1280: \section{Definitions of the initial and ending conditions of two-body models}
1281: \label{sec4.1}
1282: 
1283: As clear from the comparison of the left and right panel of Fig.~\ref{fig:eomega}, 
1284: because of the bad behaviour of the PN Hamiltonian near LSSO, we propose 
1285: as our best bet, for describing in a physically reliable manner the non-adiabatic
1286: evolution of BH binaries, and their transition between inspiral
1287: and plunge, to use an EOB-resummed Hamiltonian, we shall,
1288: for more generality, consider, and compare, in this Section several types of two-body models.
1289: 
1290: To define a specific model we must make  various choices:
1291: (i) choice of a PN-expanded (or ``Taylor-expanded'') Hamiltonian
1292: (say ``TH'') versus an EOB-resummed Hamiltonian (say ``EH'');
1293: (ii) choice of a Taylor-expanded flux function (say ``TF'')
1294: versus a Pad\'e-resummed one (say ``PF''), and finally,
1295: (iii) choice of the PN accuracies used both in the Hamiltonian
1296:  (say $ n $ PN) and the flux function (say $m$ PN). This leads
1297:  to models denoted, for instance,${\rm THTF}(n, m)$, ${\rm EHTF}(n ,m)$, ${\rm EHPF}(n, m)$.
1298:  In addition, as we are here mainly considering
1299:  the evolution of {\it spinning} binaries, we shall add
1300:  an initial letter S to recall that fact. This leads to
1301:  models denoted as ${\rm STHTF}(n, m)$,..., ${\rm SEHPF}(n, m)$.
1302:  To simplify, we shall only consider the PN accuracies $(2, 2.5)$ or $(3,3.5)$.
1303:  To further simplify, we shall focus on comparing ``fully Taylor'' models 
1304: (i.e., ${\rm STHTF}$), to ``fully resummed '' ones
1305: (i.e., ${\rm SEHPF}$). Finally, this leads us to considering
1306: only four models: ${\rm STHTF(2, 2.5)}$, ${\rm STHTF}(3, 3.5)$, ${\rm SEHPF}(2, 2.5)$, and 
1307: ${\rm SEHPF}(3, 3.5)$.
1308: [Note, as discussed above, that because of the appearance of spurious poles 
1309: in a few tests, we applied in those cases, the Pade resummation only to 
1310: the non-spinning part of the flux.]
1311: 
1312: An important parameter in our present model building is to
1313: decide when to stop the evolution. This issue was already
1314: tackled in Ref.~\cite{BD2}. There, because we were using
1315: an EOB Hamiltonian, and were considering non-spinning BH's,
1316: we found that we could follow the ``plunge'' (after LSO crossing)
1317: down to a (Schwarzschild-like) radius $\simeq 3 M$, at which
1318: point we could match to a ring-down signal made of least-damped
1319: quasi-normal modes. It was found in \cite{BD2} that,
1320: contrary to what the usually employed word ``plunge'' suggests,
1321: the inspiral motion after crossing the LSO was
1322: staying ``quasi-circular'', with a kinetic energy in the
1323: radial motion staying small
1324: in absolute value, and smaller than $0.3$ times the
1325: kinetic energy in the azimuthal motion down to $ R \simeq 3 M$.
1326: In our ``spinning'' evolutions the situation is more complicated
1327: (notably when considering large aligned spins, and also,
1328: for evident reasons, when considering Taylor-expanded Hamiltonians).
1329: We leave to future work a detailed discussion of the
1330: matching to ring-down. We decided to stop
1331: the evolution  as soon as one of the following inequalities
1332: ceased to be fulfilled:
1333: 
1334: \begin{subequations}
1335: \begin{eqnarray}
1336: %\label{cond2}
1337: %\omega &>& \omega_{\rm LSSO}^{\rm EOB}\,,\\
1338: \label{cond3}
1339: |\dot{R}| &<& 0.3 |\mathbf{V}_t|\,, \\
1340: \label{cond4}
1341: P_R^2/B(R) &<& 0.3 P^2_\phi/R^2\,, \\
1342: \label{cond5}
1343: |\dot{E}_{RR}|  &>& 0.1 |\dot{E}_{RR}^{\rm Newt}|\,,\\
1344: \label{cond6}
1345: R&>& \alpha M \,,
1346: \end{eqnarray}
1347: \end{subequations}
1348:  where $B(R) = D(R)/A(R)$ [see Eqs.~(\ref{coeffPA}), (\ref{coeffPA2})].
1349: Criteria \eqref{cond3}--\eqref{cond4} ensure that the evolution does
1350: not extend too much beyond circularity, on which our formulation for
1351: radiation-reaction force relies.
1352: The quantity $\mathbf{V}_t$ is the tangential velocity
1353: (i.e., orthogonal to the relative separation vector $\mathbf{X}$).  
1354: Criterion~\eqref{cond5} is used to
1355: avoid going into regimes where the GW energy flux goes to zero (e.g.,
1356: for Taylor-expanded flux at 2.5PN order). Criterion \eqref{cond6}
1357: (in which $\alpha \sim 1$ when using the ADM-coordinate Taylor-expanded
1358: Hamiltonian, and $\alpha \sim 2$ when using the Schwarzschild-like
1359: EOB Hamiltonian)
1360: terminates the evolution at a very small radius, in case all of the
1361: above criteria fail to take effect.
1362: 
1363: In all  cases, the
1364: instantaneous GW frequency at the time when the integration is stopped
1365: defines the \emph{ending frequency} for these waveforms. We shall
1366: also consider below extended waveforms obtained by matching
1367: a ring-down signal when this ending frequency is reached.
1368: 
1369: 
1370: \subsection{Initial conditions: quasi-spherical orbits}
1371: \label{sec4.2}
1372: 
1373: [In this section we shall use natural units $c=1=G$ and set $M=1$.]
1374: 
1375: In absence of radiation reaction (RR), spherical orbits with constant
1376: radius and orbital frequency exists under spin-orbit interactions, but
1377: cease to exist when spin-spin interactions are present (except in
1378: special situations when the spins and the orbital angular momentum are
1379: all aligned/anti-aligned). When radiation reaction is treated
1380: adiabatically, an initial spherical orbit will evolve into a sequence
1381: of spherical orbits, due to Eq.~\eqref{n11}. 
1382: In this section, we formulate a prescription to construct
1383: initial conditions for non-adiabatic evolutions, which lead to
1384: quasi-spherical orbits under spin-orbit interaction.
1385: 
1386: With spin terms kept only up to the spin-orbit order, the Hamiltonian
1387: can be re-written into a simpler form,
1388: \begin{equation}
1389: H(R,P_R,L,\chi_L)=H_{\rm no\,spin}(R,P_R,L)+2\frac{L\chi_L}{R^3}\,.
1390: \end{equation}
1391: 
1392: Here $H_{\rm no\,spin}$ are terms in the Hamiltonian that do not involve spins, and
1393: \begin{equation}
1394: L\equiv |\mathbf{L}| \,, \quad \chi_L \equiv  \mathbf{S}_{\rm eff} \cdot\hat{\mathbf{L}}.
1395: \end{equation}
1396: In this form, the Hamiltonian depends on four quantites,
1397: $\{R,P_R,L,\chi_L\}$, in which $L$ and $\chi_L$ both depend on
1398: $\{\theta,\phi,P_\theta,P_\phi\}$, while $\chi_L$ also depends on the
1399: spins. In absence of radiation reaction, the conditions 
1400: for spherical orbits written in terms of partial derivatives (indicated by a subscript $i$) with respect 
1401: to the four independent variables $\{R,P_R,L,\chi_L\}$, read
1402: \begin{eqnarray}
1403: \label{eq:circ1}
1404: \big[\dot{R}\big]_0=0 &\Rightarrow&  \left[P_R\right]_0=0\,, \\
1405: \label{eq:circ2}
1406: \big[\dot{P}_R\big]_0= 0 &\Rightarrow &\left[\left(\frac{\partial H}{\partial R}\right)_c\right]_0 =\left[\left(\frac{\partial H}{\partial R}\right)_i\right]_0 =0\,.
1407: \end{eqnarray}
1408: [Here the subscript $c$ indicates canonical partial derivatives. In
1409: the rest of this section, we shall continue to use $i$ and $c$ to
1410: distinguish between these two types of partial derivatives.] With $L$
1411: and $\chi_L$ being conserved quantities, conditions \eqref{eq:circ1}
1412: and \eqref{eq:circ2} will remain satisfied if they are initially
1413: satisfied --- which proves the existence of spherical orbits.
1414: 
1415: We now construct initial conditions for spherical orbits, {\it in
1416: absence of radiation reaction}, based on Eqs.~\eqref{eq:circ1} and
1417: \eqref{eq:circ2}.  In numerical evolutions, given a source coordinate
1418: frame, $\{\mathbf{e}_x,\mathbf{e}_y,\mathbf{e}_z\}$, we specify spherical 
1419: orbits with the following initial {\it kinetic
1420: parameters}: the orbital frequency $\omega_0$, the orbital orientation (i.e., 
1421: the normal direction to the orbital plane $[\hat{\mathbf{L}}_{\rm N}]_0 = (\vX \times 
1422: \dot{\vX})/|\vX \times \dot{\vX}|$), the spins 
1423: $[\mathbf{S}_{1,2}]_0$, and the direction of initial orbital separation
1424: $\mathbf{N}=\vX/|\vX|$, which can in turn be given by an initial orbital
1425: phase $\phi_{\rm orb}$, calculated with respect to the reference direction of
1426: $[\mathbf{S}_{\rm tot} \times\hat{\mathbf{L}}_{\rm N}]_0$,
1427: %
1428: \beq
1429: \mathbf{N}_0 = \frac{[\mathbf{S}_{\rm tot} \times\hat{\mathbf{L}}_{\rm N}]_0}{|[\mathbf{S}_{\rm tot} \times\hat{\mathbf{L}}_{\rm N}]_0|}\cos\phi_{\rm orb} +
1430:  \frac{[\hat{\mathbf{L}}_{\rm N}]_0\times[\mathbf{S}_{\rm tot} \times\hat{\mathbf{L}}_{\rm N}]_0}{|[\mathbf{S}_{\rm tot} \times\hat{\mathbf{L}}_{\rm N}]_0|}\sin\phi_{\rm orb}\,.
1431: \eeq  
1432: %
1433: We calculate initial
1434: values for $\{\mathbf{X},\mathbf{P}\}$ in three steps:
1435: \begin{enumerate} 
1436: %
1437: \item
1438: We first apply a rotation $\mathcal{R}$ such that 
1439: $[\hat{\mathbf{L}}_{\rm N}]_0 \rightarrow \mathbf{e}_z$ and 
1440: $\mathbf{N}_0 \rightarrow \mathbf{e}_x$.
1441: %
1442: %
1443: \item In spherical polar coordinates, the above step implies 
1444: %
1445: \beq
1446: \label{init:norr1}
1447: \phi_0=0\,,\;\;\;\; \theta_0=\pi/2\,.
1448: \eeq
1449: %
1450: [The $\phi_0$ here should not to be confused with the orbital phase
1451: $\phi_{\rm orb}$ above.] Then, we specify the initial frequency
1452: $\omega_0$ and impose
1453: \begin{eqnarray}
1454: \label{init:norr2}
1455: \omega_0 = \dot\phi_0=\left[\left(\frac{\partial H}{\partial P_\phi}\right)_c\right]_0\,,&& 
1456:  0=\dot\theta_0 =\left[\left(\frac{\partial H}{\partial P_\theta} \right)_c\right]_0;\\
1457: \label{init:norr3}
1458: \left[P_R\right]_0=0\,, && \left[\left(\frac{\partial H}{\partial R}\right)_c\right]_0=0\,,
1459: \end{eqnarray}
1460: and solve for the four variables $\{R,P_R,P_\theta,P_\phi\}_0$. \item
1461: Finally, we apply the inverse rotation $\mathcal{R}^{-1}$ to the
1462: entire system, obtaining a set of initial spherical-orbit
1463: conditions consistent with the specified initial kinetic parameters.
1464: \end{enumerate}
1465: 
1466: When radiation reaction is included, we proceed as in Ref.~\cite{BD2} and modify Eq.~\eqref{eq:circ1} 
1467: to include a non-zero $\dot{R}$, according to the prediction from adiabatic evolution, 
1468: \begin{equation}
1469: \label{RdotfromdEdR}
1470: \big[\dot{R}\big]_0= \left[\frac{\dot{E}_{\rm RR} }{(dE/dR)_{\rm sph}}\right]_0\,,
1471: \end{equation}
1472: in order to prevent radial oscillations. [The subscript sph in Eq.~(\ref{RdotfromdEdR}) and 
1473: below denote quantities evaluated along spherical orbits.]
1474: Equations~\eqref{eq:circ2} can be kept unchanged, since $\dot{P}_R$ is second order in radiation
1475: reaction.  We now calculate $(dE/dR)_{\rm sph}$ in terms of the
1476: simplified set of independent variables,
1477: $\{R,P_R,L,\chi_L\}$. Consider neighboring spherical orbits in an
1478: adiabatic sequence, we have
1479: \begin{equation}
1480: dH = \left(\frac{\partial H}{\partial R}\right)_i dR
1481: +
1482:  \left(\frac{\partial H}{\partial P_R}\right)_i dP_R
1483:  +\left(\frac{\partial H}{\partial L}\right)_i dL
1484:  +\left(\frac{\partial H}{\partial \chi_L}\right)_i d\chi_L \,, 
1485:  \end{equation}
1486:  \begin{equation}
1487: \left(\frac{\partial H }{\partial R}\right)_i = 0,\;\;
1488: d\left(\frac{\partial H }{\partial R}\right)_i =0, \;\;
1489: P_R=0,\;\; dP_R=0\,.
1490:  \end{equation}
1491: It is straightforward to deduce that
1492: \begin{equation}
1493: \label{dEdRfull}
1494: \left(\frac{dE}{dR}\right)_{\rm sph} =
1495: -\frac{\displaystyle \bigg(\frac{\partial H}{\partial L}\bigg)_i\bigg(\frac{\partial^2 H}{\partial R^2}\bigg)_i}
1496: {\displaystyle \bigg(\frac{\partial^2 H}{\partial R\partial L}\bigg)_i}
1497: +
1498: \underbrace{
1499: \left[
1500: \left(\frac{\partial H}{\partial \chi_L}\right)_i
1501: -
1502: \frac
1503: {\displaystyle \bigg(\frac{\partial H}{\partial L}\bigg)_i
1504: \bigg(\frac{\partial^2 H}{\partial R \partial \chi_L}\bigg)_i
1505: }
1506: {\displaystyle \bigg(\frac{\partial^2 H}{\partial R\partial L}\bigg)_i}
1507: \right]
1508: \left(
1509: \frac{d\chi_L}{d R}
1510: \right)_{\rm sph}}_{\mbox{will be ignored}}\,.
1511: \end{equation}
1512: The second term on the right-hand side can be ignored, as we argue
1513: later in this section, because $\chi_L$ is still conserved to a high
1514: accuracy even in presence of radiation reaction. In special
1515: configurations with $\hat{\mathbf{L}} =\mathbf{e}_z$ (or equivalently
1516: $\theta=\pi/2$, $P_\theta=0$) we can re-write Eq.~\eqref{dEdRfull} in
1517: terms of canonical variables in spherical-polar coordinates:
1518: \begin{equation}
1519: \label{dEdRsimple}
1520: \left(\frac{dE}{dR}\right)_{\rm sph}
1521: =
1522: -\left[\frac{(\partial H/\partial P_\phi)_c (\partial^2 H/\partial R^2)_c}
1523: {(\partial^2 H / \partial R \partial P_\phi)_c}\right]_{\theta=\pi/2,P_\theta=0}
1524: \,.
1525: \end{equation}
1526: We also note that when $\hat{\mathbf{L}}$ is known to be $\mathbf{e}_z$, we can 
1527: calculate $\dot{E}_{\rm RR}$, from Eq.~\eqref{n13} right away using only $\omega_0$ and $[\mathbf{S}_{1,2}]_0$.
1528: 
1529: We can now construct quasi-spherical initial conditions when radiation
1530: reaction is present. As done in Ref.~\cite{BD2}, up to leading order in 
1531: radiation reaction, we only need to augment our no-radiation-reaction 
1532: initial conditions with a {\it non-zero} $P_R$, with initial values for all other canonical
1533: variables unchanged. In order to do so, we {\it insert} three more
1534: steps between steps 2 and 3 above:
1535: \begin{itemize}
1536: \item[2a] During step 2, we have obtained a set of spherical-polar-coordinate initial conditions, for a {\it rotated system} with $\{\mathbf{N}_0,[\hat{\mathbf{L}}_{\rm N}]_0\}=\{\mathbf{e}_x,\mathbf{e}_z\}$.  The canonical angular momentum, $[\mathbf{L}]_0$ of this system, though, will not in general be along $\mathbf{e}_z$. However, 
1537: being orthogonal to $\mathbf{N}_0=\mathbf{e}_x$, it must be within the $\mathbf{e}_y-\mathbf{e}_z$ plane.  We now apply a further rotation $\mathcal{R}'$ around $\mathbf{N}_0 = \mathbf{e}_x$ to the entire system, such that afterwards  
1538: we have $\{\mathbf{N}_0,[\hat{\mathbf{L}}]_0\}=\{\mathbf{e}_x,\mathbf{e}_z\}$, i.e., $\theta_0=\pi/2$ and $[P_\theta]_0=0$.
1539: %
1540: \item[2b]  Now that Eq.~\eqref{dEdRsimple} is applicable and $\dot{E}_{\rm RR}$ is readily obtainable, we insert them into Eq.~\eqref{RdotfromdEdR} and obtain $[\dot{R}]_0$. [Note that in this process we use the set of initial conditions already obtained for a spherical orbit in absence of radiation reaction, with $P_R=0$.] From this $[\dot{R}]_0$ , we obtain the initial value of $[P_R]_0$ to insert into our existing set of initial conditions:
1541: \begin{equation}
1542: \left[P_R\right]_0 = \frac{[\dot{R}]_0}{\displaystyle \left[\frac{1}{P_R}\left(\frac{\partial H}{\partial P_R}\right)_c\right]_{P_R \rightarrow 0}}\,.
1543: \end{equation}
1544: \item[2c] Gathering our new set of $\{R,\theta,\phi,P_R,P_\theta,P_\phi\}_0$, we obtain the Cartesian-coordinate variables, and apply a inverse rotation $(\mathcal{R}')^{-1}$ to the entire system. [Now again we have $\{N_0,[\hat{\mathbf{L}}_{\rm N}]_0\}=\{\mathbf{e}_x,\mathbf{e}_z\}$, and are ready to proceed to step 3.]
1545: \end{itemize}
1546: 
1547: \comment{
1548: We impose our initial conditions in a spherical polar coordinate system. 
1549: We place our initial orbital angular momentum pointing to the
1550: $+z$ direction, and the initial separation to be along the $+x$ direction, i.e., 
1551: %
1552: \beq
1553: \label{init:thetaphi}
1554: \theta_0=\pi/2 \,,\quad (P_\theta)_0=0\,,\quad \phi_0=0\,.
1555: \eeq
1556: %
1557: we also require the initial orbital frequency to be $\omega_0$:
1558: %
1559: \beq
1560: \label{init:phidot}
1561: \omega_0 \equiv \sqrt{\dot\phi_0^2+\dot\theta_0^2} =\sqrt{
1562: \left(\frac{ \partial H}{\partial P_\phi}\right)_0^2+
1563: \left(\frac{ \partial H}{\partial P_\theta}\right)_0^2}; 
1564: \qquad \dot\phi_0=\left(\frac{\partial H}{\partial P_\phi}\right)_0>0\,.
1565: \eeq
1566: %
1567: [Note that in absence of spins, $P_\theta=0$ gives $\dot\theta_0=0$, 
1568: so the orbital plane points to the $+z$ direction and $\omega_0=\dot\phi_0$.]  
1569: In absence of radiation reaction, in order to obtain orbits with constant radius 
1570: ({\em spherical orbits}), we need to have $\dot{R}_0=(\partial H/\partial P_R)_0=0$, 
1571: which requires 
1572: %
1573: \begin{equation}
1574: \label{init:PR}
1575: (P_R)_0=0\,,
1576: \end{equation}
1577: %
1578: since the non-spinning part of the conservative Hamiltonian, $H_{n \rm PN}$, 
1579: is quadratic in $P_R$, and the spinning parts are independent
1580: from $P_R$. In order to maintain a constant radius, we must
1581: further require 
1582: %
1583: \beq
1584: \label{init:dPR}
1585: \left (\frac{d P_R}{dt} \right )_0 = - \left ( \frac{\partial H}{\partial R} 
1586: \right )_0 =0\,.
1587: \eeq
1588: %
1589: If spin-spin effects ($H_{SS}$) are ignored, spherical orbits are
1590: compatible with the Hamiltonian, namely, the initial conditions
1591: \eqref{init:PR} and \eqref{init:dPR} will guarantee a constant radius 
1592: throughout the evolution (while the orbital orientation and spin
1593: directions will change); if spin-spin effects are included, circular
1594: orbits in general are {\it not} compatible with the Hamiltonian, and
1595: conditions  \eqref{init:PR} and \eqref{init:dPR} will give orbits with
1596: oscillating orbital radius. However, the orbits they provide should be
1597: as close as possible to spherical orbits. Together, the $6$
1598: equations, \eqref{init:thetaphi}--\eqref{init:dPR}, completely specify
1599: the $6$ orbital variables, $(R,\theta,\phi,P_R,P_\theta,P_\phi)$; the
1600: two initial spin vectors are specified separately. 
1601: 
1602: With radiation reaction present in our evolution, the radius will
1603: shrink at the radiation reaction time scale. In order to be as close
1604: as possible to an adiabatic sequence of  shrinking spherical orbits, 
1605: we need to augment Eq.~\eqref{init:PR} with the adiabatic rate of radius change: 
1606: %
1607: \begin{equation}
1608: \label{init:PR:RR}
1609: \left(\frac{\partial H}{\partial P_R}\right)_0= (\dot{R})_{\rm spher}
1610: =\frac{(dE/dt)_0}{\left[(dH/dR)_{\rm spher}\right]_0}\,.
1611: \end{equation}
1612: %
1613: The quantity $\left[(dH/dR)_{\rm spher}\right]_0$ is the change of energy between 
1614: neighboring spherical orbits and we will now determine it by assuming that 
1615: initial conditions are given when the two bodies are so far apart that 
1616: spin-spin couplings can be neglected.
1617: 
1618: The energy of a spherical orbit can be written {\it only}  
1619: as function of $R$ and $\mathbf{S}_{\rm eff} \cdot \hat{\mathbf{L}}$, 
1620: i.e., $H_{\rm spher}(R,\mathbf{S}_{\rm eff} \cdot \hat{\mathbf{L}})$. 
1621: To obtain this result we use the fact that the total Hamiltonian, including 
1622: only spin-orbit terms, can be written as (using $P^2 = P^2_R+ L^2/R^2$)
1623: %
1624: %
1625: where $L = |\mathbf{L}|$ includes dependence on $\theta$, $\phi$,
1626: $P_\theta$ and $P_\phi$. As said, spherical orbits require $P_R=0$ and 
1627: $\partial H/\partial R=0$, so for any pair of neighboring spherical 
1628: orbits we have
1629: %
1630: \beq
1631: \label{eq:H:L}
1632: d H = \frac{\partial H }{\partial L}\,dL + 
1633: \frac{\partial H }{\partial (\mathbf{S}_{\rm eff} \cdot \hat{\mathbf{L}})}
1634: \,d (\mathbf{S}_{\rm eff} \cdot \hat{\mathbf{L}})\,.
1635: \eeq
1636: %
1637: By imposing $d(\partial H/\partial R)=0$ we obtain 
1638: %
1639: \begin{equation}
1640: \label{eq:dHdr}
1641: 0 = \frac{\partial^2 H}{\partial R^2} dR + \frac{\partial^2 H}{\partial R\, \partial L} d L + 
1642: \frac{\partial^2 H}{\partial R \partial (\mathbf{S}_{\rm eff} \cdot \hat{\mathbf{L}})} d 
1643: (\mathbf{S}_{\rm eff} \cdot \hat{\mathbf{L}})  
1644: \,.
1645: \end{equation}
1646: %
1647: As we shall show below, $\mathbf{S}_{\rm eff} \cdot \hat{\mathbf{L}}$ is 
1648: conserved with high accuracy throughout the adiabatic evolution, so we 
1649: neglect its variation in Eqs.~\eqref{eq:H:L} and \eqref{eq:dHdr}, and obtain
1650: %
1651: \begin{equation}
1652: \label{init:dHdR}
1653: \left(\frac{dH}{dR}\right)_{\rm spher} = 
1654: -\left[\frac{(\partial
1655:     H/\partial L)(\partial^2 H/\partial R^2)}{(\partial^2 H
1656:     /\partial R\,\partial L)}\right]_{P_R=0,\, (\partial H/\partial
1657:   R)=0}\,, 
1658: \end{equation}
1659: %
1660: which should be plugged in Eq.~(\ref{init:PR:RR}).
1661: }
1662: 
1663: A straightforward analysis of the various error terms allows us 
1664: to conculde that the fractional error of assuming that  $\chi_L \equiv \mathbf{S}_{\rm
1665: eff}\cdot \hat{\mathbf{L}} $ is constant along the adiabatic evolution is 
1666: of 3PN order. 
1667: 
1668: We note that our steps 1, 2, (2a, 2b, 2c), and 3 can still be applied
1669: to give initial conditions, even if the Hamiltonian contains spin-spin
1670: terms, although the orbits that follow will in general have
1671: oscillatory radius and orbital frequency, due to the non-existence of
1672: quasi-spherical orbits.  In Fig.~\ref{fig:rdot} we show the evolutions
1673: of $\dot{r}/(r\omega)$ with (dark curves) and without (light curves)
1674: spin-spin terms, for $(10+10)M_{\odot}$ (left panel) and
1675: $(15+5)M_{\odot}$ (right panel) binaries. We start evolution at
1676: 40\,Hz, with 
1677: $(\theta_{S_1},\phi_{S_1}\theta_{S_2},\phi_{S_2})=(60^\circ,90^\circ;60^\circ,0^\circ)$,
1678: and show the evolution up to 200\,Hz.
1679: 
1680: 
1681: \section{Comparison of waveforms and evaluation of overlaps}
1682: \label{sec4.3}
1683: 
1684: In harmonic coordinates, the gravitational wave emitted by a binary system at the 
1685: leading quadrupole order, in terms of metric perturbation at a distance $D$, is
1686: %
1687: \beq
1688: h_{i j} = \frac{H_{ij}}{D} \equiv \frac{2\mu}{D}\,\frac{d^2}{d t^2} (X_i\,X_j)\,.
1689: \eeq
1690: Using the leading-order equation of motion, $\ddot{X}_k = - 
1691: {M\,X_k}/{R^3}$, we re-write the normalized perturbation $H_{ij}$ as: 
1692: %
1693: \beq
1694: \label{hGW0}
1695: H_{i j} = {4\mu}\,\left (V_i\,V_j -M \frac{X_i\,X_j}{R^3} \right )\,.
1696: \eeq
1697: %
1698: Here $X_i$ and $V_i \equiv \dot{X}_i$ can be obtained straightforwardly
1699: by solving the Hamilton equations. 
1700: Depending on the wave propagation direction and the orientation of the
1701: detector, the metric perturbation $h_{ij}$ has to be contracted with
1702: an appropriate ``detection tensor" to give the actually detected
1703: waveform. For this we refer, for instance, to Sec.~IIIC of Ref.~\cite{pbcv1} and
1704: Sec.~II of Ref.~\cite{bcv2} (in particular see Eq.(15)). 
1705: 
1706: Following Ref.~\cite{bcv2}, the parameters in precessing binaries can be 
1707: distinguished in {\it binary local parameters} $\{m_1, m_2, S_1, S_2,
1708: \theta_{\rm S1}, \theta_{\rm S2}, \phi_{\rm S1} - \phi_{\rm S2}\}$, 
1709: {\it binary directional parameters} $\{ \theta_{\rm L}, \phi_{\rm L}, \phi_{\rm S1} + \phi_{\rm S2}\}$ 
1710: (which determine the orientation of the binary as a whole in space) 
1711: and {\it directional parameters} $\{\Theta, \varphi, \theta, \phi, \psi \}$, 
1712: describing the GW-propagation and the detector orientation 
1713: [see Table I in Ref.~\cite{bcv2} and discussion around it].
1714: To these parameters we need to add the initial time and the initial orbital phase. 
1715: 
1716: In the {\it precessing} convention introduced in Ref.~\cite{bcv2}, 
1717: the GW signal can be neatly written in terms of: (i) parameters 
1718: depending on the observer's location and orientation, 
1719: $\{\Theta, \varphi, \theta, \phi, \psi \}$, which are time 
1720: independent, initial time and initial orbital phase (henceforth denoted 
1721: as extrinsic parameters) and (ii) parameters depending on the details 
1722: of the dynamics, $\{m_1, m_2, S_1, S_2, \theta_{\rm S1}, \theta_{\rm S2}, 
1723: \phi_{\rm S1} - \phi_{\rm S2}\}$ (henceforth denoted as intrinsic 
1724: parameters). The GW signal does not depend on the binary directional 
1725: parameters, $\{ \theta_{\rm L}, \phi_{\rm L}, \phi_{\rm S1} + \phi_{\rm S2}\}$, 
1726: since those parameters can be re-absorbed in the definition of the source 
1727: frame at initial time and in the directional parameters $\{\Theta, 
1728: \varphi, \theta, \phi, \psi \}$ through a rigid rotation of the detector-binary system. 
1729: 
1730: The distinction between extrinsic and intrinsic parameters is due originally
1731: to Sathyaprakash and Owen~\cite{Sathya,O}. Extrinsic parameters are parameters 
1732: which change the signal shape in such a way that we do not actually need to 
1733: lay down templates in the bank along those parameter directions, saving computational time. 
1734: By contrast we need to lay down templates along the directions 
1735: of the intrinsic parameters. In Refs.~\cite{bcv2, pbcv1},
1736: a semi-analytical method to maximize over the extrinsic 
1737: parameters in precessing binaries has been proposed.
1738: 
1739: In view of the bad performances of the Taylor-expanded Hamiltonian\footnote{
1740: We have in mind here the absence of LSSO. 
1741: Recall also that when the binary mass ratio is significantly
1742: different from one, one can firmly conclude that the Taylor-expanded
1743: Hamiltonian is a poor representation of the dynamics, while the
1744: EOB-resummed one is  a good one.}, we a priori expect that the
1745: waveforms computed from STHTF models will be significantly
1746: different from the SEHPF ones. It remains, however, interesting
1747: to {\it measure their difference} in the data-analysis sense,
1748: i.e.\ to compute the {\it overlaps} between the two types
1749: of waveforms. If it happened that, after maximization
1750: over all possible parameters, the overlap between the two
1751: types of signals were very close to unity, one could still
1752: consider the Taylor models as {\it effectual} (in the
1753: sense of \cite{DIS98}) representations
1754: of the EOB models. However, for practical reasons, we
1755: did not try to embark on a full maximization of the overlaps.
1756: For simplicity, we {\it only} tackled the maximization over the
1757:  extrinsic parameters, and {\it not} on the intrinsic
1758: ones. The resulting partially maximized overlap is therefore
1759: only a lower bound of the fully maximized overlap. Still, this
1760: result can be considered as a reasonable measure of  the ``closeness'' between the
1761: two sorts of models (especially because we do not wish to
1762: use models which would have significantly different physical
1763: parameters).
1764: 
1765: 
1766: \subsection{Lack of ``closeness'' between Taylor and Effective-One-Body models}
1767: 
1768: In Tables~\ref{overlapmax3pn} and \ref{overlapmax2pn}  
1769: we study the {\it closeness} (in the sense just defined of overlap
1770: maximized only over the extrinsinc parameters\footnote{More precisely,
1771: we do the maximization over the extrinsic parameters of the EOB model. 
1772: Though this introduces an asymmetry in the definition of the {\it closeness},
1773: we do not expect this asymmetry to be physically significant.})
1774: between STHTF(3,3.5) and SEHPF(3,3.5), as well
1775: as  between STHTF(2,2)~\footnote{We use SHT(2,2) instead of STHTF(2,2.5) 
1776: because for equal-mass binaries
1777: the Taylor-approximant for the flux at 2.5PN order
1778: becomes negative for large values of $v$~\cite{DIS98}},
1779: and SEHPF(2,2.5) models.
1780: 
1781: We consider three typical binary masses $(10 + 10) M_\odot$,
1782: $(15 + 15) M_\odot$ and $(15 + 5) M_\odot$, and several initial 
1783: spin orientations~\footnote{For these data we always refer the initial spin 
1784: directions to the initial direction of the orbital 
1785: Newtonian angular momentum, as specified in Fig. 4 of Ref.~\cite{bcv2},  
1786: and we set the initial direction of the Newtonian orbital angular 
1787: momentum along the $x$-axis of the source frame (i.e., we fix 
1788: $\theta_{\rm L} = \pi/2$ and $\phi_{\rm L}=0$, see Fig. 3 in Ref.~\cite{bcv2}).}.
1789: We always fix the pattern functions $F_+ = 1$, $F_\times = 0$ and GW
1790: propagation parameters $\Theta = \pi/4$ and $\varphi = 
1791: 0$ [for notations and definitions see Sec.~IIIC of Ref.~\cite{pbcv1}
1792: and Sec.~II of Ref.~\cite{bcv2}]. 
1793: The initial frequency is always set to $f_{\rm in} = 30\,{\rm Hz}$ and 
1794: the ending frequency $f_{\rm end}$ is determined by one of the 
1795: criteria in Eqs.~\eqref{cond3}--\eqref{cond6}.
1796: In Tables~\ref{overlapmax3pn} and \ref{overlapmax2pn} the two black holes 
1797: are assumed to carry maximal and half-maximal spins, respectively. 
1798: We list the ending frequency, the LSSO frequency  and the BH radial 
1799: separation at $t_{\rm fin}$ for the template and target, 
1800: together with two types of overlaps: the overlaps maximized over the initial
1801: time and initial orbital phase only ($\rho_{\rm max, 2}$),
1802: and the overlaps maximized over those parameters and
1803: $\{\Theta, \varphi, \alpha = f(\theta, \phi, \psi) \}$,
1804: as well, ($\rho_{\rm max, 5}$), using the semi-analytical 
1805: method suggested in Ref.~\cite{pbcv1}.
1806: Table~\ref{overlapmax3pn} and ~\ref{overlapmax2pn} also contains the non-spinning case. 
1807: 
1808: As these tables show, the two types of models are not
1809: at all ``close to each other''. The overlaps are indeed
1810: quite low, as low as $0.32$. The overlaps evidently increase when
1811: we maximize over five rather than two extrinsic parameters, but
1812: not dramatically, and only for binaries with high and comparable mass, 
1813: with initial spins lying in the half-space 
1814: opposite (with respect to the orbital plane) 
1815: to the initial orbital angular momentum. In this case 
1816: the dynamical evolution is shorter, since the LSSO occurs 
1817: at lower frequency [see also Figs. \ref{fig:eomegaEOB}, \ref{fig:e-a} ], 
1818: and the differences in  STHTF(3,3.5) and SEHPF(3,3.5) can be compensated by an offset in the
1819: extrinsic parameters of the template with respect to the target. 
1820: Moreover, both the conservative dynamics for circular orbits 
1821: and the GW flux predicted by SEHPF-approximants and STHTF-approximants, 
1822: are closer in the anti-aligned case than in the aligned case, 
1823: as can be see in Figs.~\ref{Fig6} and \ref{Fig7}. 
1824: 
1825: However, when the binary mass ratio is significantly
1826:  different from 1, the number
1827: of modulational cycles increases, and the differences in the 
1828: two models can no longer be compensated by re-adjusting the 
1829: template extrinsic parameters. When the initial spins are lying 
1830: in the same half-space of the orbital angular momentum, the evolution is longer, the LSSO 
1831: happens at high frequency [see also Figs. \ref{fig:eomegaEOB}, \ref{fig:e-a} ], 
1832: and in this case, even for high, comparable masses the differences both in 
1833: the conservative and non-conservative late dymanics in the two models cannot 
1834: be compensated by a bias in the template extrinsic parameters. 
1835: 
1836: {}From Tables~\ref{overlapmax2pn} we 
1837: notice that all the above considerations apply also at 2PN order,  
1838: where the differences in the conservative 
1839: and non-conservative dynamics of STHTF and SEHPF approximants are
1840: even larger. We checked that these considerations do not change much 
1841: when spins are smaller, say half-maximal. 
1842: 
1843: Having confirmed that Taylor models cannot be considered
1844: as being effectively close to the EOB ones, we shall
1845: only use in the following the a priori better EOB models.
1846: 
1847: \subsection{Negligible influence of the of the ``transverse component'' of the
1848: radiation reaction force.}
1849: 
1850: Having in mind possible simplifications of the models,
1851: we first investigated the relevance of the second term
1852: in the R.H.S. of Eq.~(\ref{n18}), i.e. the component of
1853: the radiation reaction force which is ``transverse'',
1854: in the sense of being directed along $L$, and
1855: therefore orthogonal to the main ``longitudinal term'',
1856: which is parallel to the momentum $\vP$).
1857: In Table~\ref{overlapnoFL} we study the influence of this transverse
1858: component of the RR force on the dynamics and the waveforms.
1859: For the binary masses $(10 + 10) M_\odot$, $(15 + 5) M_\odot$ 
1860: and a few initial spin orientations, we evaluate the same 
1861: quantities of Table~\ref{overlapmax3pn}, when including and not 
1862: including the RR force along $L$ [see second and third term in Eq.~(\ref{n18})]. 
1863: We give here only the overlaps maximized over five extrinsic parameters.
1864: We find that $\rho_{\rm max, 5}$ is larger than $\sim 0.98$
1865: in all cases. We therefore conclude that it would
1866: suffice to use a simplified RR force parallel to the
1867: linear momentum $\vP$ (as in the non-spinning circular case).
1868: One should, however, include, for better accuracy,
1869:  in the coefficient of $P_i$
1870: in Eq.~(\ref{n18}) the spin-dependent terms.
1871: 
1872: \subsection{Influence of the resummation of the ``longitudinal'' part of
1873: the radiation reaction}
1874: 
1875: We consider here the ``longitudinal'' part of the radiation reaction,
1876: i.e. the first term on the R.H.S. of Eq.~(\ref{n18}). This term is
1877: given by the flux function, which was written in Eq.~(\ref{n13}) above
1878: as a straightforward PN-expansion.  One can therefore either leave
1879: this longitudinal component in non-resummed, ``Taylor'' form, or
1880: choose to resum it by means of Pad\'e approximants.  In
1881: Table~\ref{overlapflux} we investigate how the choice of the flux
1882: function (Taylor-expanded or Pad\'e-resummed) may affects the dynamics
1883: and the waveforms. Using in all cases the EOB Hamiltonian to describe
1884: the dynamics, we evaluate the overlaps between models using a Taylor
1885: flux and models using a Pad\'e flux. [We maximize over the five
1886: extrinsic parameters of the EOB model.]  We find that, when the initial 
1887: spins are lying in the same half-space of the orbital 
1888: angular momentum, after maximization over the five extrinsic parameters, 
1889: the overlaps are reasonably large (larger than $0.84$), but still lower
1890: than unity. We obtain much higher overlaps when the initial spins are not lying 
1891: in the same half-space of the orbital angular momentum. These results 
1892: are consistent with Figs.~\ref{Fig5} and Fig.~\ref{Fig7}. 
1893: 
1894: If we assume that the equal-mass flux function is a smooth deformation 
1895: of the test-mass limit one, since previous findings \cite{DIS98,BD2,DIS01,PS} 
1896: in the test-mass limit case pointed out the usefulness of Pad\'e-resumming 
1897: the flux function, we would conclude that Pad\'e-resummed fluxes are better
1898: approximants of the numerically determined flux also in th equal-mass case.
1899: 
1900: \subsection{Negligible influence of the quadrupole-monopole terms}
1901: 
1902: Still in the spirit of trying to simplify the models to their crucial elements,
1903: Table~\ref{overlapnoQM} investigates how waveforms are
1904: affected by the quadrupole-monopole terms, and Table~\ref{overlapadiab} 
1905: studies how the evolution obtained by averaging the spin terms over a 
1906: period may differ from the non-adiabatic evolution. 
1907: Considering the high values of $\rho_{\rm max, 5}$ we obtain
1908: in both cases, we can say that the quadrupole-monopole interaction and the adiabaticity 
1909: of the spin terms, have little physical effects over the dynamics and waveforms.
1910: The differences can be compensated by re-adjusting the template extrinsic parameters.
1911: 
1912: \subsection{Influence of the initial orbital phase}
1913: 
1914: Finally, we investigated the influence of the initial orbital phase
1915: (all other quantities being fixed) on the waveform.
1916: In an adiabatic evolution in which spin terms are averaged over a 
1917: period the joint evolution of $\hat{\mathbf{L}}_{\rm N}$ and $\mathbf{S}$ is not affected by the
1918: initial orbital phase. As a consequence, two configurations with the
1919: same initial values for $\hat{\mathbf{L}}_{\rm N}$ and $\mathbf{S}$,
1920: but different orbital phases $\phi_{\rm orb}$ will keep the difference
1921: between orbital phases unchanged through the evolution. This may not
1922: be true in our non-adiabatic evolution for two reasons: (i) the spin-spin
1923: interaction Hamiltonian depends explicitly on the separation vector
1924: $\mathbf{N}$ [see Eq.~(\ref{3.3})], and (ii) the evolution 
1925: depends on the canonical orbital angular momentum,
1926: which is not orthogonal to the orbital plane. We illustrate this 
1927: feature by evolving two maximally spinning
1928: $(15+15)M_\odot$ binaries with initial orbital phases (at 40\,Hz)
1929: differing by $\pi/2$, and all other parameters identical:
1930: $(\theta_{S_1},\phi_{S_1};\theta_{S_2},\phi_{S_2})=(60^\circ,90^\circ;60^\circ,0^\circ)$. In
1931: Fig.~\ref{fig:rel-orb-phase}, we plot the difference $\Delta\phi_{\rm
1932: orb}$ between their relative orbital phases measured with respect to
1933: $\hat{\mathbf{L}}_N$. This difference grows in time, and accumulates
1934: around $270^\circ$ by the end of the evolution.  We also show
1935: waveforms detected with $(F_+,F_\times;\Theta,\varphi)= (1,0;\pi/4,0)$
1936: in Fig.~\ref{fig:orbphase}.  Their phases start out to differ by
1937: $180^{\circ}$ as expected, and non-adiabatic effects drive them away
1938: by more than 2 cycles toward the end of the evolution. We notice that 
1939: comparing the waveform is less straightforward than comparing the
1940: relative orbital phase, because the waveform phase can differ from
1941: twice the orbital phase, due to precessions.
1942: 
1943: \section{Losses of energy and angular-momentum and the waveform including ringdown}
1944: \label{sec4.4}
1945: 
1946: In the following, we use as model the ``best bet'' we can make, i.e.
1947: the spinning EOB Hamiltonian\footnote{We did not investigate the
1948: ``closeness'' between the models derived from the spinning
1949: Hamiltonian used here, and those deduced from
1950: the further resummed, Kerr-like EOB Hamiltonian proposed in \cite{TD}.
1951: In view of the comparison showed in Fig.~\ref{fig:e-a},
1952: we expect that the two models are very close to each other.} 
1953: with a Pad\'e-resummed flux. Both being taken 
1954: to the highest PN-accuracy available, i.e. $n=3, m=3.5$, in the
1955: notation used above.
1956: 
1957: In Ref.~\cite{BD01}, using the non-spinning EOB Hamiltonian at 2PN order, it was 
1958: found that the energy emitted during the plunge is $ \sim 0.7 \%$ of $M$, 
1959: with a comparable energy loss $ \sim  0.7 \% $ of $ M$ during the ring-down phase.
1960: This gives a total energy released beyond the LSSO in the non-spinning case of 
1961: $\sim 1.4\%$ of $M$ to be contrasted with $4-5\%$ of $M$ 
1962: estimated in Ref.~\cite{BBCLT}, where the authors use a combination of 
1963: numerical and perturbative approximation methods. Note also that
1964:  Flanagan and
1965: Hughes~\cite{FH98} predicted $\sim 10 \% M$ for inspiral and plunge, 
1966: and $\sim 3 \% M$ for ring-down phase. 
1967: 
1968: Here, to have more confidence in our EOB-based estimates,
1969: we decided to use {\it three different ways  } of evaluating
1970:  the energy radiated in the spinning case.
1971: We used, at once, (i) the change, along the evolution (between
1972: some initial frequency and some final one) in the numerical value
1973: of the Hamiltonian (\ref{Hspineob}) ($\delta E_H$),  (ii) the time-integral
1974: of the square of the third derivative of the
1975:  quadrupole moment $I_{ij}$ with $i,j=1,2,3$, i.e.
1976: %
1977: \beq
1978: \frac{d E_I}{dt} = \frac{1}{5} \frac{d^3 I_{i j}}{dt^3}\,\frac{d^3 I_{i j}}{dt^3}
1979: \quad \quad I_{ij} = \mu\,\left (X_i\,X_j - \frac{1}{3}\delta_{ij}X^k\,X_k\right)\,.
1980: \eeq
1981: %   
1982: and (iii) the time-integral of the
1983: energy flux carried away by our leading-order quadrupole waveform,
1984: %
1985: \beq
1986: \frac{d E_h}{dt} = \frac{1}{20}\,\int \sum_{i j} \dot{H}_{i j}^{\rm TF} \dot{H}_{ij}^{\rm TF}\,,
1987: \eeq
1988: %
1989: with $H_{ij}^{\rm TF}$ being
1990: the trace-free part of the normalized metric perturbation,
1991: $H_{ij}$ [see Eq.~(\ref{hGW0}) above]. 
1992: 
1993: 
1994: In Fig.~\ref{fig:comp:eloss} we compare the accumulated energy release
1995: from these three prescriptions, for a $(15+15)M_{\odot}$ maximally
1996: spinning binary with a generic set of spin orientations when starting
1997: evolution at $f_{\rm GW}=30\,$Hz:
1998: $(\theta_{S_1},\phi_{S_1};\theta_{S_2},\phi_{S_2})=(60^\circ,90^\circ;60^\circ,0^\circ)$. In
1999: the left panel, we keep radiation reaction force at the Newtonian
2000: order, while we use 3.5PN Pad\'e flux in the right panel. As we see
2001: from the figure, these prescriptions differ more from each other when
2002: 3.5PN radiation reaction is used instead of Newtonian --- this is
2003: consistent with the fact that both $\delta E_I$ and $\delta E_h$
2004: involve quadrupole radiation only; furthermore, for lower frequencies
2005: the $\delta E_H$ curve lies below those of $\delta E_I$ and $\delta
2006: E_h$, which is consistent with the fact that Post-Newtonian GW
2007: luminosity is in general smaller than the Newtonian prediction. The
2008: difference among $\delta E_I$ and $\delta E_h$ can be attributed to
2009: the difference between PN (in our case EOB at 3PN) and Newtonian
2010: dynamics, which seems to be small till around $f_{\rm
2011: GW}=200$\,Hz in our case (which corresponds to $v\approx 0.45$).
2012: 
2013: 
2014: The rather satisfactory agreement between the various ways of
2015: estimating the energy loss gives us some confidence in our
2016: EOB-based estimates. In the following, we shall use the a priori
2017: best estimate (because it is the one which involves the highest
2018: PN accuracy): the one based on the change in the total
2019: EOB  Hamiltonian $H$. [We use  Pad\'e-resummed fluxes,
2020: and the two combined PN-accuracies $(n,m) = (2,2.5)$
2021: and $(n,m) = (3,3.5)$.]
2022: 
2023: In the upper panels of Fig.~\ref{fig:eloss} we plot
2024: the accumulative energy loss $\delta E_H$
2025: starting from $f_{\rm GW}=40\,$Hz as a
2026: function of the instantaneous GW frequency $f$, for $(15+15)M_{\odot}$
2027: (upper left panel) and $(15+5)M_{\odot}$ (upper right panel) binaries,
2028: each for 4 sets of initial spin orientations with {\it maximal spins}:
2029: aligned (dash-dot-dot curves), antialigned (dotted curves),
2030: $(\theta_{S_1},\phi_{S_1};\theta_{S_2},\phi_{S_2})=(60^\circ,90^\circ;60^\circ,0^\circ)$
2031: (denoted by {\it generic-up}, dash-dot curves), and
2032: $(\theta_{S_1},\phi_{S_1};\theta_{S_2},\phi_{S_2})=(120^\circ,90^\circ;120^\circ,0^\circ)$
2033: (denoted by {\it generic-down}, dashed curves), as well as for the
2034: non-spinning configuration (solid curves).  We use both SEHPF(3,3.5)
2035: (dark curves) and SEHPF(2,2.5) (light curves) models. In each panel, we
2036: also use vertical grid lines to mark LSSO frequencies. For the
2037: SEHPF(2,2.5) model, all our evolutions go beyond their corresponding
2038: LSSO frequencies. The situation is a bit different for the SEHPF(3,3.5).
2039: Indeed, as was shown in \cite{TD} and in Fig.~\ref{fig:e-a} above,
2040: the 3PN-EOB LSSO for mostly aligned fast-spinning BH's is drastically
2041: drawn inwards towards very high orbital frequencies.
2042: So high, indeed, that, for aligned and generic-up configurations,
2043: they fall out of the frequency range plotted in Fig.~\ref{fig:eloss}.
2044: As a consequence, for aligned and generic-up configurations, the dynamical
2045: evolutions become rather non-adiabatic even before the formal LSSOs
2046: is reached.
2047: 
2048: As we see from the plots, within a given GW frequency interval, binaries tend to
2049: emit more energy in configurations where spins are more aligned with
2050: the orbital angular momentum. This agrees with the results
2051: of \cite{TD} and of Fig.~\ref{fig:eomega} above, showing
2052: that more aligned configurations are drawn towards more deeply bound
2053: states. This, together with the fact that LSSO frequencies are
2054: pushed higher in aligned configurations (as we also see from
2055: Fig.~\ref{fig:eomega}), can make the total energy releases in aligned
2056: configurations several times more than those in anti-aligned
2057: configurations. In Table \ref{tab:EJ}, we list values of $\delta
2058: E_H/M$, accumulated from $40$\,Hz up to LSSO frequency (if reached)
2059: or ending frequency, otherwise, for configurations plotted in
2060: Fig.~\ref{fig:eloss}.  We also list the energy released below 40\,Hz,
2061: and values of energy release when 2PN Hamiltonian and 2.5PN Pad\'e
2062: flux are used.  For $(15+15)M_{\odot}$, maximally spinning binaries,
2063: the energy released from 40\,Hz up to the end of our evolution (determined 
2064: by one of the criteria \eqref{cond3}--\eqref{cond6})
2065: can range from $0.6\%$ of $M$ in the (anti-aligned configuration) to
2066: $5\%$ of $M$ (anti-aligned configuration), with the non-spinning
2067: configuration releasing $1.6\%\sim 1.8\%$ of $M$ (in which $0.8\% \sim
2068: 1.1\%$ of $M$ is released before LSSO).  For $(15+5)M_{\odot}$
2069: binaries, the range is similar, from $0.5\%$ to $5\%$ of $M$, with
2070: non-spinning configuration releasing $1.2\% \sim 1.4\%$ of M (in which
2071: $0.7\%\sim 0.8\%$ of M from before the LSSO).  We also note that the
2072: energies of around $ 0.8\%$ of M and $0.5\%$ of M are released below
2073: 40\,Hz, for $(15+15)M_\odot$ and $(15+5)M_\odot$ binaries,
2074: respectively. See also Eq. (4.1) of \cite{TD} for an approximate
2075: analytical estimate of the energy released down to the LSSO,
2076: as a function of both $\eta = m_1 m_2/M^2$ and $\chi_L$.
2077: 
2078: \subsection{Evolution of the dimensionless rotation parameter $J/E^2$}
2079: 
2080: An important consistency check of the EOB approach concerns
2081: the dimensionless total angular momentum
2082:  ratio $|\mathbf{J}|/E^2$ ultimately reached
2083: by spinning black hole binaries. Indeed, if the EOB method would,
2084: at the end of its validity domain, predict a ratio $|\mathbf{J}|/E^2$
2085: larger than unity, this would preclude to match this end
2086: state with the newly born Kerr black hole expected
2087: from the coalescence of the two initial (spinning) black holes.
2088: This issue was investigated {\it in the adiabatic} approximation
2089: in \cite{TD}. There, it was shown that, when using the
2090: 3PN-accurate EOB Hamiltonian, the ratio $|\mathbf{J}|/E^2$
2091: estimated {\it at the LSSO}, was {\it always smaller than unity}.
2092: This result was not at all guaranteed in advance, and resulted
2093: from a delicate  competition between the linear increase of $|\mathbf{J}|$
2094: when increasing the (aligned) spin of the individual BH's,
2095: and its non-linear decrease because of the displacement of the LSSO
2096: towards smaller radii for aligned spinning configurations.
2097: It was found in \cite{TD} that the maximum of
2098: $(|\mathbf{J}|/E^2)_{\rm LSSO}$ was about $0.83$, and was reached for
2099: $\chi_L \simeq + 0.3$.
2100: 
2101: The fact that this maximum value is significantly below one,
2102: leaves room for not running into any consistency problem even when
2103: taking into account the further changes of both $E$ and $\vJ$
2104: during the plunge that follows the crossing of the LSSO.
2105: Though our present attack on the problem does not properly
2106: consider the final matching between the plunge and the
2107: formation of a final Kerr hole, it goes beyond the previous
2108: treatments in going beyond the adiabatic approximation.
2109: In the lower panels of Fig.~\ref{fig:eloss}, we plot the
2110: continuous time evolution of the ratio $|\mathbf{J}|/E^2$
2111: during the late stages of the inspiral and its subsequent
2112: non-adiabatic ending (which, in many cases, except
2113: in fact for the most dangerous  aligned, is a
2114: post-LSSO plunge). It is convenient to use
2115: the gravitational wave frequency $f_{\rm GW}$ to label
2116: the ``time'' along this evolution. Satisfactorily, we
2117:  observe that, for all the binaries we have considered,
2118: $|\mathbf{J}|/E^2$ decrease to below $1$ before either the
2119: LSSO or the end of our evolution, whichever comes first. This means
2120: there are no a priori obstacles to having a Kerr
2121: black hole form right after the end of the non-adiabatic ``quasi-plunge''.
2122: This means also that, contrary to an early
2123: suggestion \cite{FH98}
2124: based on rather coarse estimates, there is no ground for expecting
2125:  a large emission of gravitational waves between the plunge
2126:  and the merger. In Table~\ref{tab:EJ} we also list the values
2127: of $|\mathbf{J}|/E^2$, at LSSO frequency (if reached) and ending
2128: frequency, for SEP(2,2.5) and SEP(3,3.5) models.
2129: 
2130: In a pioneering work,
2131: Baker et al.~\cite{BCLT} evaluated by a 3D numerical
2132: simulation the energy radiated from moderately spinning
2133: BH binaries with spins aligned or anti-aligned with the orbital angular-momentum.
2134: They started the (very short) numerical evolution close to the LSSO predicted 
2135: by the effective potential method of Pfeiffer et al.~\cite{PTC}.
2136: As already mentioned above, the numerical initial data chosen in these
2137: works rely only on an Initial Value Problem (IVP)
2138:  formulation, and significantly differ both from the numerical initial data
2139: constructed by the HKV method \cite{GGB, Cook} and from
2140: the predictions made by the EOB method (while the HKV
2141: and EOB results are quite close to each other \cite{DGG,CookPfeiffer}).
2142: For instance, \cite{DGG} estimated that the ratio between the
2143: orbital frequencies $\omega_{\rm IVP}/\omega_{\rm EOB}$ was about 2.
2144: One should probably wait until HKV-type initial data for
2145: spinning BH's are evolved until coalescence to meaningfully
2146: compare their results with the results derived above for
2147: energy releases within the EOB approach. However,
2148: to have an idea of the current distance between analytical
2149: estimates and numerical ones, we
2150:  have determined the energy released between the LSSO and
2151: the final frequency, at 2PN and 3PN order, for two of the spin configurations 
2152: investigated by Baker et al. For spins aligned (anti-aligned) and  
2153: $\chi_1 = \chi_2 = 0.17$ ($\chi_1 = \chi_2 = 0.25$), we find that 
2154: the energy released is $\sim (0.6 - 0.9) \% M$ [$\sim (1 - 3)\% M$]. 
2155: Baker et al. found $\sim (1.7-1.9) \% M$ and $\sim (1.9-2.1) \% M$, 
2156: respectively. It should be noted that the energy released 
2157: evaluated by Baker et al. includes also the ring-down phase. Ours does not.
2158: 
2159: \subsection{Complete waveforms describing the  non-adiabatic inspiral and coalescence
2160: of precessing binary black holes}
2161: 
2162: Finally, though we have not yet carefully studied at which stage
2163: we could meaningfully join our ``quasi-plunge'' evolution to the
2164: formation of a ringing BH, we have decided, to show the promise
2165: of a purely analytical EOB-base approach to follow
2166:  Ref.~\cite{BD2} in matching (by
2167: requiring first-order continuity of the emitted waveform)
2168:  the end of our  waveform (here defined by the first violation
2169:  of the ``adiabatic criteria'' \eqref{cond3}--\eqref{cond4}) to
2170:  a ringdown waveform generated from the lowest $l=m=2$
2171: quasi-normal mode of a Kerr black hole. We determine
2172: the  mass and spin parameters of the final hole
2173:  by the energy and angular momentum of the binary at the
2174: end of our evolution:
2175: %
2176: \begin{equation}
2177: M_{\rm BH} = E_{\rm fin}\,,\quad a_{\rm BH} = \left[{|\mathbf{J}|}/{E^2}\right]_{\rm fin}\,.
2178: \end{equation}
2179: %
2180: In Fig.~\ref{fig:ringdown}, we plot the complete waveforms,
2181: so obtained,  for non-spinning, and
2182: maximally spinning $(15+15)M_\odot$ binaries in the generic-up and
2183: generic-down configurations (these refer to the configuration at
2184: $f_{\rm GW}=40\,$Hz, the starting point of evolution). We have shifted
2185: these waveforms in time so that the end of inspiral evolutions all
2186: happen at $t=0$.  Notice that at this stage the waveform which includes 
2187: the ring-down phase should be considered as an example.  
2188: Indeed, by restricting ourselves to the quasi-normal mode $l=m=2$, 
2189: we have tacitly assumed that the total angular momentum at the time 
2190: the ring-down phase starts is dominated by the orbital angular momentum. 
2191: However, this is not generally the case when spins are present and 
2192: the quasi-normal modes with $l\neq 2$ might be excited, as well, 
2193: and contribute to the waveform. A more thourough analysis is left for 
2194: the future.
2195: 
2196: 
2197: \section{Conclusions}
2198: \label{sec5}
2199: 
2200: We provided a first attack on the problem of analytically determining
2201: the gravitational waveforms emitted during the last stages of
2202: dynamical evolution of precessing binaries of spinning black holes,
2203: i.e. during the non-adiabatic ending of the inspiral phase, and its
2204: transition to a plunge. We reviewed the various available Hamiltonian
2205: descriptions of the dynamics of spinning black hole (BH) binaries, and
2206: studied (following \cite{TD}) the characteristics of the stable
2207: spherical orbits that exist when spin-spin effects are neglected
2208: compared to spin-orbit ones. We derived the contribution to radiation
2209: reaction (for quasi-circular orbits) which is linear in the spins. Our
2210: results agree with the corresponding recent results of \cite{Will}. We then
2211: used this analytical description of the radiation-reaction-driven
2212: inspiral of spinning binaries to construct non-adiabatic models of
2213: coalescing binary waveforms.  We compared the various models and
2214: concluded, in confirmation of previous results, that our current
2215: ``best bet'' for a non-adiabatic model describing the transition from 
2216: adiabatic inspiral to plunge is obtained by combining: (i) an effective one body (EOB)~
2217: \cite{BD1,BD2}, 3PN-accurate \cite{DJS} resummed Hamiltonian, including 
2218: spin-dependent interactions \cite{TD}, with (ii) Pad\'e-resummed radiation reaction force 
2219: (including spin-terms). Conclusion (i) is rather robust, since 
2220: as Fig.~\ref{fig:eomega} shows, the PN-expanded Hamiltonian does not show any 
2221: LSSO and differs significantly from the PN-expanded analytically computed 
2222: function ${E}(\Omega)$; conclusion (ii) is based on the assumption 
2223: that the flux function in the equal-mass case is a smooth deformation of 
2224: the test-mass limit result. Since in the latter case Pad\'e approximants 
2225: were shown \cite{DIS98,PS} to have better agreement with exact 
2226: numerical flux functions, we would conclude that this is also true in the 
2227: equal-mass case. 
2228: 
2229: Our main results, obtained by means of this ``best bet'' EOB model are:
2230: 
2231: (1) An estimate of the energy and angular momentum released by the
2232: binary system during its last stages of evolution:  inspiral, transition
2233: from inspiral to plunge, and plunge;
2234: 
2235: (2) The finding (which confirms the conclusions of \cite{TD}) that
2236: the dimensionless rotation parameter $j/E^2$ is always smaller than unity
2237: at the end of the inspiral;
2238: 
2239: (3) The construction of complete waveforms, approximately describing
2240: the entire gravitational-wave emission process from
2241: precessing binaries of spinning black holes: adiabatic inspiral,
2242: non-adiabatic transition between inspiral and plunge, plunge,
2243: merger and ringdown. Following \cite{BD2} these waveforms were
2244: constructed by matching a quasi-normal-mode ringdown to the end
2245: of the plunge signal. These tentative complete waveforms are
2246:  preliminary because we did not include here a careful study
2247: of how to join, in a physically motivated manner, the last stages
2248: of the plunge to the merger phase. They extend, however, the
2249: (better justified) complete waveforms constructed in \cite{BD2}
2250: to the more genral case of spinning and precessing binaries.
2251: 
2252: \acknowledgements
2253: 
2254: A.B. thanks the Max-Planck Institut f\"ur Gravitationsphysik (Albert-Einstein-Institut) for support during her visit.  Y.C.'s research is supported by Alexander von Humboldt Foundation's Sofja Kovalevskaja Award (funded by the German Federal Ministry of Education and Research), and by the NSF grant PHY-0099568 (during his stay at Caltech); he also thanks the Institut d'Astrophysique de Paris (CNRS) for support during his visit.
2255: 
2256: \begin{thebibliography}{999}
2257: \frenchspacing
2258: %
2259: \bibitem{LIGO} A. Abramovici et al., Science {\bf 256}, 325 (1992); 
2260: \url{http://www.ligo.caltech.edu}.
2261: %
2262: \bibitem{VIRGO} B. Caron et al., Class. Quantum Grav. {\bf 14}, 1461 (1997); 
2263: \url{http://www.virgo.infn.it}.
2264: %
2265: \bibitem{GEO} H. L\"uck et al., Class. Quantum Grav. {\bf 14}, 1471 (1997); 
2266: \url{http://www.geo600.uni-hannover.de}.
2267: %
2268: \bibitem{TAMA} M. Ando et al., Phys. Rev. Lett. {\bf 86}, 3950 (2001); 
2269: \url{http://tamago.mtk.nao.ac.jp}.
2270: %
2271: \bibitem{TD} T.~Damour, Phys.\ Rev.\ D {\bf 64}, 124013 (2001).
2272: %
2273: \bibitem{BD1} A. Buonanno and T. Damour, Phys. Rev. D {\bf 59}, 084006 (1999). 
2274: %
2275: \bibitem{BD2} A. Buonanno and T. Damour, Phys. Rev. D {\bf 62}, 064015 (2000).
2276: %
2277: \bibitem{ACST94} T. A. Apostolatos, C. Cutler, G.J. Sussman and K.S. Thorne, {Phys. Rev. D} {\bf 49}, 6274 (1994).
2278: %
2279: \bibitem{K} L. E. Kidder, Phys. Rev. D {\bf 52}, 821 (1995).
2280: %
2281: \bibitem{apostolatos0} T. A. Apostolatos, {Phys. Rev. D} {\bf 52}, 605 (1995).
2282: %
2283: \bibitem{apostolatos1} T. A. Apostolatos, {Phys. Rev. D} {\bf 54}, 2421 (1996).
2284: %
2285: \bibitem{apostolatos2} T. A. Apostolatos, {Phys. Rev. D} {\bf 54}, 2438 (1996).
2286: %
2287: \bibitem{GKV} P. Grandcl\'ement, V. Kalogera  and A. Vecchio, {Phys. Rev. D} {\bf 67}, 042003 (2003). 
2288: %
2289: \bibitem{GK} P. Grandcl\'ement and V. Kalogera, {Phys. Rev. D} {\bf 67}, 082002 (2003). 
2290: %
2291: \bibitem{bcv2} A. Buonanno, Y. Chen, and M. Vallisneri, {Phys. Rev. D} \textbf{67}, 104025 (2003).
2292: %
2293: \bibitem{pbcv1} Y. Pan, A. Buonanno, Y. Chen and M. Vallisneri, Phys. Rev. D {\bf 69}, 104017 (2004).
2294: %
2295: \bibitem{bcpv1} A. Buonanno, Y. Chen, Y. Pan and M. Vallisneri, Phys. Rev. D {\bf 70}, 104003 (2004). 
2296: %
2297: \bibitem{Gpc} P. Grandcl\'ement, M. Ihm, V. Kalogera, and K. Belczynski, Phys. Rev. D {\bf 69}, 102002 (2004). 
2298: %
2299: \bibitem{DIS98} T. Damour, B.R. Iyer and B.S. Sathyaprakash, Phys. Rev. D {\bf 57}, 885 (1998). 
2300: %
2301: \bibitem{DIS00}T.~Damour, B.~R.~Iyer and B.~S.~Sathyaprakash, Phys.\ Rev.\ D {\bf 62}, 084036 (2000).
2302: %
2303: \bibitem{DJS} T. Damour, P. Jaranowski and G. Sch\"afer, Phys. Rev. D {\bf 62}, 084011 (2000).
2304: %
2305: \bibitem{DIS01} T. Damour, B.R. Iyer and B.S. Sathyaprakash, 
2306: Phys. Rev. D {\bf 63}, 044023 (2001); {\it ibidem} {\bf 66}, 027502 (2002).
2307: %
2308: \bibitem{DGG} T. Damour, E. Gourgoulhon and P. Grandcl\'ement, Phys. Rev. D {\bf 66}, 024007 (2002).
2309: 
2310: \bibitem{GGB} P. Grandcl\'ement, E. Gourgoulhon and S. Bonazzola, Phys. Rev. D {\bf 65}, 044021 (2002).
2311: %
2312: \bibitem{bcv1} A. Buonanno, Y. Chen, and M. Vallisneri, {Phys. Rev. D} \textbf{67}, 024016 (2003).
2313: %
2314: \bibitem{DIS03} T.~Damour, B.~R.~Iyer, P.~Jaranowski and B.~S.~Sathyaprakash, 
2315: Phys.\ Rev.\ D {\bf 67}, 064028 (2003).
2316: %
2317: \bibitem{BBCLT} J. Baker, B. Br\"ugmann, M. Campanelli, C.O. Lousto and R. Takahashi,
2318: Phys. Rev. Lett. {\bf 87}, 121103 (2001).
2319: %
2320: \bibitem{BCLT} J. Baker, M. Campanelli, C.O. Lousto and R. Takahashi,
2321: Phys. Rev. D {\bf 69}, 027505 (2004).
2322: %
2323: \bibitem{K00} V. Kalogera, Astrophys. J. {\bf 541}, 319 (2000).
2324: %
2325: \bibitem{DS88} T. Damour and G. Sch\"afer, Nuov. Cimento {\bf 101} (1988) 127.
2326: %
2327: \bibitem{JS98} P.~Jaranowski and G.~Schafer,
2328:   Phys.\ Rev.\ D {\bf 57}, 7274 (1998)
2329:   [Erratum-ibid.\ D {\bf 63}, 029902 (2001)].
2330: %
2331: \bibitem{DJSd} T.~Damour, P.~Jaranowski and G.~Schafer, Phys.\ Lett.\ B {\bf 513}, 147 (2001).
2332: %
2333: \bibitem{EP} E. Poisson, Phys. Rev. D {\bf 57} (1998) 5287.
2334: %
2335: \bibitem{Will} C.~M.~Will, Phys.\ Rev.\ D {\bf 71}, 084027 (2005).
2336: %
2337: \bibitem{DJSPoincare}T.~Damour, P.~Jaranowski and G.~Schafer,
2338:   Phys.\ Rev.\ D {\bf 62}, 021501 (2000)
2339:   [Erratum-ibid.\ D {\bf 63}, 029903 (2001)].
2340: %
2341: \bibitem{KWW}
2342:  L.~E.~Kidder, C.~M.~Will and A.~G.~Wiseman,
2343:   Phys.\ Rev.\ D {\bf 47}, 3281 (1993).
2344: %
2345: \bibitem{WS} N.~Wex and G.~Schafer, Class.\ Quantum Grav.
2346: {\bf 10}, 2729 (1993).
2347: %
2348: \bibitem{BGH} B.M. Barker, S.N. Gupta and R.D. Haracz, Phys. Rev. {\bf 149} (1966) 1027.
2349: %
2350: \bibitem{BO} B.M. Barker and R.F. O'Connell, Phys. Rev. D {\bf 2} (1970) 1428; 
2351: B.M. Barker and R.F. O'Connell, Gen. Relativ. and Gravit. 
2352: {\bf 11} (1979) 149; {\bf 5} (1974) 539.
2353: %
2354: \bibitem{TH} K.S. Thorne and J.B. Hartle, Phys. Rev. D {\bf 31} (1985) 1815.
2355: %
2356: 
2357: \bibitem{PTC} H. Pfeiffer, S.A. Teukolsky and G.B. Cook, Phys. Rev. D 62, 104018 (2000).
2358: %
2359: \bibitem{Cook} G.~B.~Cook, Phys.\ Rev.\ D {\bf 65}, 084003 (2002).
2360: %
2361: \bibitem{CookPfeiffer} G.~B.~Cook and H.~P.~Pfeiffer,
2362:   Phys.\ Rev.\ D {\bf 70}, 104016 (2004).
2363: %
2364: \bibitem{LB} L. Blanchet, Phys. Rev. D {\bf 65}, 124009 (2002).
2365: %
2366: \bibitem{DJSinvar} T.~Damour, P.~Jaranowski and G.~Schafer,
2367:   Phys.\ Rev.\ D {\bf 62}, 044024 (2000).
2368: %
2369: \bibitem{BDE}L.~Blanchet, T.~Damour and G.~Esposito-Farese,
2370:   Phys.\ Rev.\ D {\bf 69}, 124007 (2004).
2371: %
2372: \bibitem{Itoh} Y.~Itoh,  Phys.\ Rev.\ D {\bf 69}, 064018 (2004).
2373: %
2374: \bibitem{DJS01} T. Damour, P. Jaranowski and G. Sch\"afer, Phys. Rev. D {\bf 63}, 044021 (2001).
2375: %
2376: \bibitem{ABF01} V. de Andrade, L. Blanchet and G. Faye, Class. Quantum Grav. {\bf 18}, 753 (2001).
2377: %
2378: \bibitem{BIJ02} L. Blanchet, B.R. Iyer and B. Joguet, Phys. Rev. D {\bf 65}, 064005 (2002).
2379: %
2380: \bibitem{B98} L. Blanchet, Class. Quantum Grav. {\bf 15}, 113 (1998).
2381: %
2382: \bibitem{BDEI04} L. Blanchet, T. Damour, G. Esposito-Far\`ese and B.R. Iyer, Phys. Rev. Lett. {\bf 93}, 091101 (2004).
2383: %
2384: \bibitem{fluxnoS} E. Poisson, Phys. Rev. D {\bf 52}, 5719 (1995).
2385: %
2386: \bibitem{fluxS} M. Shibata (private communication).
2387: %
2388: \bibitem{PS} E. Porter and B.S. Sathyaprakash, Phys. Rev. D {\bf 71}, 024017(2005)
2389: %
2390: \bibitem{Sathya} B. S. Sathyaprakash, Phys. Rev. D \textbf{50}, R7111 (1994).
2391: %
2392: \bibitem{O} B. J. Owen, \prd \textbf{53}, 6749--6761 (1996).
2393: %
2394: \bibitem{BD01} A. Buonanno and T. Damour , contributed paper to the IX$^{\rm th}$ Marcel 
2395: Grossmann Meeting (Rome, July 2000); 2000, gr-qc/0011052.
2396: %
2397: \bibitem{FH98} E.E. Flanagan and S.A. Hughes, Phys. Rev. D {\bf 57}, 4535 (1998).
2398: \end{thebibliography}
2399: 
2400: \begin{table}
2401: \begin{tabular}{c|cccccccc}
2402: \hline
2403: \hline
2404: $(\theta_{\rm S1}, \phi_{\rm S2}, \theta_{\rm S1}, \phi_{\rm S2}) $
2405:   & $f_{\rm fin}^{\rm SEP}$ (Hz) & $f_{\rm LSSO}^{\rm SEP}$(Hz) & 
2406: ${\rm R}_{\rm fin}^{\rm SEP}/M$ & $f_{\rm fin}^{\rm SHT}$(Hz) & 
2407: $f_{\rm LSSO}^{\rm SHT}$(Hz) &
2408: ${\rm R}_{\rm fin}^{\rm SHT}/M$ 
2409: & $\rho_{\rm max, 2}$ & $\rho_{\rm max, 5}$ \\
2410: \hline
2411: \hline
2412: \multicolumn{9}{l}{\hspace{6cm}$(10+10)M_\odot$}\\
2413: \hline
2414: \hline
2415: no spin & 289 & 285 & 4.8 & 287 & 285 & 4.2 & 0.9150 & - \\
2416: \hline
2417: $(0^o, 0^o, 0^o, 0^o)$ & 745  & 2145  & 2.5 & 466 & 2145 & 2.8 & 0.3750 &  -   \\	
2418: $(180^o, 0^o, 0^o, 0^o)$  & 290  & 285  & 4.8 & 285 & 285 & 4.2 & 0.9166 &  -   \\	
2419: $(180^o, 0^o, 180^o, 0^o)$  & 145  & 145 & 8.1 & 145 & 145 & 7.4 & 0.5587 &  -   \\	
2420: $(60^o, 90^o, 60^o, 0^o)$ & 633 & 873 & 2.6 & 502 & 502 & 2.6 & 0.4851 & 0.5883 \\ 
2421: $(120^o, 90^o, 60^o, 0^o)$ & 280 & 280 & 4.9 & 269 & 269 & 4.4 & 0.5420 & 0.9472 \\ 
2422: $(120^o, 90^o, 120^o, 0^o)$ & 187 & 187 & 6.7 & 186 & 186 & 6.0 & 0.6096 & 0.9536 \\ 
2423: \hline
2424: \hline
2425: \multicolumn{9}{l}{\hspace{6cm}$(15+15)M_\odot$} \\
2426: \hline
2427: \hline
2428: no spin & 192 & 190 & 4.8 & 192 & 190 & 4.2 & 0.8137 & - \\
2429: \hline
2430: $(0^o, 0^o, 0^o, 0^o)$ & 497  & 1430  & 2.6 & 311 & 1430 & 2.6 & 0.4550 &  -   \\	
2431: $(180^o, 0^o, 0^o, 0^o)$  & 193  & 190  & 4.8 & 191 & 190 & 4.2 & 0.8148 &  -   \\	
2432: $(180^o, 0^o, 180^o, 0^o)$  & 98  & 97 & 8.0 & 97 & 97 & 7.4 & 0.6403 &  -   \\	
2433: $(60^o, 90^o, 60^o, 0^o)$ & 429 & 758 & 2.6 & 347 & 347 & 2.4 & 0.5067 & 0.6290  \\ 
2434: $(120^o, 90^o, 60^o, 0^o)$ & 185 & 185 & 4.9 & 180 & 180 & 4.4 & 0.5112 & 0.9456  \\ 
2435: $(120^o, 90^o, 120^o, 0^o)$ & 124 & 124 & 6.7 & 124 & 124 & 6.7 & 0.6868 & 0.9684 \\ 
2436: \hline
2437: \hline
2438: \multicolumn{9}{l}{\hspace{6cm}$(15+5)M_\odot$}\\
2439: \hline
2440: \hline
2441: no spin & 267 & 265 & 5.2 & 265 & 265  &4.2 & 0.6023 & - \\
2442: \hline
2443: $(0^o, 0^o, 0^o, 0^o)$ & 862  & 1442  & 2.3 & 479 & 1442 & 2.5 & 0.3268 &  -   \\	
2444: $(180^o, 0^o, 0^o, 0^o)$  & 176  & 175  & 7.0 & 177 & 175 & 6.2 & 0.5188 &  -   \\	
2445: $(180^o, 0^0, 180^o, 0^o)$  & 141  & 140 & 8.3 & 140 & 140 & 7.5 & 0.4445 &  -   \\	
2446: $(60^o, 90^o, 60^o, 0^o)$ & 715 & 796 & 2.4 & 425 & 743 & 2.4 & 0.4478 & 0.6111  \\ 
2447: $(120^o, 90^o, 60^o, 0^o)$ & 208 & 207 & 6.2 & 208 & 208 & 5.3 & 0.5471 & 0.7496  \\ 
2448: $(120^o, 90^o, 120^o, 0^o)$ & 224 & 224 & 5.9 & 225 & 225 & 4.9 & 0.5735 & 0.8360  \\ 
2449: \hline
2450: \end{tabular}
2451: \caption{We list the overlaps between STHTF(3,3.5), used as target model, and 
2452: SEHPF(3,3.5), used as template model, for several binary masses and 
2453: initial spin orientations.  
2454: The two black holes are assumed to carry maximal spins $\chi_1=\chi_2=1$, 
2455: but for comparion we also show the results in absence of spins. 
2456: The evolution starts at $f_{\rm in} = 30$ Hz. In the first three columns 
2457: we list the ending frequency, the LSSO frequency  and the 
2458: BH radial separation at $t_{\rm fin}$ for the template model and 
2459: in the second three columns we show the same quantities but for the target 
2460: model. The last two columns contains the overlap maximized only over 2 extrinsic 
2461: parameters $\rho_{\rm max,2}$ and maximed over 5 extrinsic parameters 
2462: $\rho_{\rm max, 5}$, as described in the text. [Notice that for the spin configuration 
2463: $(60^o, 90^o, 60^o, 0^o)$ and masses $(10+10)M_\odot$ and $(15+5)M_\odot$, due to 
2464: a pole in the Pad\'e-approximant flux, we apply the Pad\'e resummation only to 
2465: the non-spinning terms in the flux.]  
2466: \label{overlapmax3pn}}
2467: \end{table}
2468: 
2469: \begin{table}
2470: \begin{tabular}{c|cccccccc}
2471: \hline
2472: \hline
2473: $(\theta_{\rm S1}, \phi_{\rm S2}, \theta_{\rm S1}, \phi_{\rm S2}) $
2474:   & $f_{\rm fin}^{\rm SEP}$ (Hz) & $f_{\rm LSSO}^{\rm SEP}$(Hz) & 
2475: ${\rm R}_{\rm fin}^{\rm SEP}/M$ & $f_{\rm fin}^{\rm SHT}$(Hz) & 
2476: $f_{\rm LSSO}^{\rm SHT}$(Hz) &
2477: ${\rm R}_{\rm fin}^{\rm SHT}/M$ 
2478: & $\rho_{\rm max, 2}$ & $\rho_{\rm max, 5}$ \\
2479: \hline
2480: \hline
2481: \multicolumn{9}{l}{\hspace{6cm}$(10+10)M_\odot$}\\
2482: \hline
2483: \hline
2484: no spin & 242 & 237 & 5.6 & 237 & 237 & 4.5 & 0.4691 & - \\
2485: \hline
2486: $(0^o, 0^o, 0^o, 0^o)$ & 628  & 628  & 2.9 & 628 & 628 & 2.6 & 0.3170 &  -   \\	
2487: $(180^o, 0^o, 0^o, 0^o)$  & 237  & 237  & 5.6 & 237 & 237 & 4.5 & 0.4681 &  -   \\	
2488: $(180^o, 0^o, 180^o, 0^o)$  & 139  & 139 & 8.4 & 139 & 139 & 7.6 & 0.6433 &  -   \\	
2489: $(60^o, 90^o, 60^o, 0^o)$ & 367 & 367 & 4.1 & 342 & 342 & 3.5 & 0.4197 & 0.4882 \\ 
2490: $(120^o, 90^o, 60^o, 0^o)$ & 234 & 234 & 5.7 & 229 & 229 & 4.7 & 0.4220 & 0.6015 \\ 
2491: $(120^o, 90^o, 120^o, 0^o)$ & 173 & 173 & 7.1 & 172 & 172 & 6.3 & 0.6681 & 0.9556 \\ 
2492: \hline
2493: \hline
2494: \multicolumn{9}{l}{\hspace{6cm}$(15+15)M_\odot$} \\
2495: \hline
2496: \hline
2497: no spin & 158 & 158 & 5.6 & 158 & 158 & 4.5 & 0.4880 & - \\
2498: \hline
2499: $(0^o, 0^o, 0^o, 0^o)$ & 419  & 419  & 2.9 & 419 & 419 & 2.6 & 0.4044 &  -   \\	
2500: $(180^o, 0^o, 0^o, 0^o)$  & 158  & 158  & 5.6 & 158 & 158 & 4.5 & 0.4885 &  -   \\	
2501: $(180^o, 0^o, 180^o, 0^o)$  & 93  & 93 & 8.4 & 93 & 93 & 7.6 & 0.7140 &  -   \\	
2502: $(60^o, 90^o, 60^o, 0^o)$ & 240 & 238 & 4.2 & 241 & 241 & 4.1 & 0.4549 & 0.5186  \\ 
2503: $(120^o, 90^o, 60^o, 0^o)$ & 156 & 155 & 5.7 & 152 & 152 & 4.7 & 0.4827 & 0.6767  \\ 
2504: $(120^o, 90^o, 120^o, 0^o)$ & 115 & 115 & 7.1 & 116 & 116 & 6.2 & 0.7227 & 0.9442 \\ 
2505: \hline
2506: \hline
2507: \multicolumn{9}{l}{\hspace{6cm}$(15+5)M_\odot$}\\
2508: \hline
2509: \hline
2510: no spin & 232 & 232 & 5.7 & 233 & 232  & 4.7 & 0.6111 & - \\
2511: \hline
2512: $(0^o, 0^o, 0^o, 0^o)$ & 608  & 608  & 2.9  & 608 & 608 & 2.6 & 0.2695 &  -   \\	
2513: $(180^o, 0^o, 0^o, 0^o)$  & 167  & 166  & 7.3 & 166 & 166 & 6.5 & 0.8720 &  -   \\	
2514: $(180^o, 0^0, 180^o, 0^o)$  & 136  & 136 & 8.5 & 136 & 136 & 7.7 & 0.6743 &  -   \\	
2515: $(60^o, 90^o, 60^o, 0^o)$ & 352 & 352 & 4.2 & 367 & 367 & 3.2 & 0.2696 & 0.4978  \\ 
2516: $(120^o, 90^o, 60^o, 0^o)$ & 192 & 192 & 6.6 & 191 & 191 & 5.7 & 0.6566 & 0.8173  \\ 
2517: $(120^o, 90^o, 120^o, 0^o)$ & 169 & 169 & 7.2 & 167 & 167 & 6.5 & 0.6207 & 0.8970  \\ 
2518: \hline
2519: \end{tabular}
2520: \caption{Overlaps between STHTF(2,2), used as target model, and 
2521: SEHPF(2,2.5), used as template model, for several binary masses and 
2522: initial spin orientations.  
2523: The two black holes are assumed to carry maximal spins $\chi_1=\chi_2=1$, 
2524: but for comparion we also show the results in absence of spins. 
2525: The evolution starts at $f_{\rm in} = 30$ Hz. In the first three columns 
2526: we list the ending frequency, the LSSO frequency  and the 
2527: BH radial separation at $t_{\rm fin}$ for the template model and 
2528: in the second three columns we show the same quantities but for the target 
2529: model. The last two columns contains the overlap maximized only over 2 extrinsic 
2530: parameters $\rho_{\rm max,2}$ and maximed over 5 extrinsic parameters 
2531: $\rho_{\rm max, 5}$, as described in the text. 
2532: \label{overlapmax2pn}}
2533: \end{table}
2534: 
2535: \begin{table}
2536: \begin{tabular}{c|ccccccc}
2537: \hline
2538: \hline
2539: $(\theta_{\rm S1}, \phi_{\rm S2}, \theta_{\rm S1}, \phi_{\rm S2}) $
2540: & $f_{\rm fin}^{\rm SEP, noF_L}$ (Hz) & 
2541: $f_{\rm LSSO}^{\rm SEP, noF_L}$ (Hz) & ${\rm R}_{\rm fin}^{\rm SEP, noF_L}/M$ 
2542: & $f_{\rm fin}^{\rm SEP}$(Hz) & $f_{\rm LSSO}^{\rm SEP}$ (Hz) &
2543: ${\rm R}_{\rm fin}^{\rm SEP}/M$ & $\rho_{\rm max, 5}$ \\
2544: \hline
2545: \hline
2546: \multicolumn{8}{l}{\hspace{6cm}$(10+10)M_\odot$}\\
2547: \hline
2548: \hline
2549: $(60^o, 90^o, 60^o, 0^o)$ & 633 & 872 & 2.7 & 660 & 1211 &2.7 & 0.9860 \\ 
2550: $(120^o, 90^o, 120^o, 0^o)$ & 187 & 186 & 6.7 & 186 & 186 &6.7 & 0.9953 \\ 
2551: \hline
2552: \hline
2553: \multicolumn{8}{l}{\hspace{6cm}$(15+5)M_\odot$} \\
2554: \hline
2555: \hline
2556: $(60^o, 90^o, 60^o, 0^o)$ & 743 & 767 & 2.3 & 564 & 564 &2.9 & 0.9839 \\ 
2557: $(120^o, 90^o, 120^o, 0^o)$ & 178 & 177 & 7.0 & 179 & 179 &6.9 & 0.9969 \\ 
2558: \hline
2559: \hline
2560: \end{tabular}
2561: \caption{Effect of radiation-reaction force along $L$ over the binary evolution and 
2562: waveforms by comparing SEHPF(3,3.5) with no $F_L$, used as target model, and 
2563: SEHPF(3,3.5), used as template model, for several binary masses and 
2564: initial spin orientations. 
2565: The two black holes are assumed to carry maximal spins $\chi_1=\chi_2=1$. 
2566: The evolution starts at $f_{\rm in} = 30$ Hz. In the first three columns 
2567: we list the ending frequency, the LSSO frequency  and the 
2568: BH radial separation at $t_{\rm fin}$ for the template model and 
2569: in the second three columns we show the same quantities but for the target 
2570: model. The last two columns contains the overlap maximized only over 2 extrinsic 
2571: parameters $\rho_{\rm max,2}$ and maximed over 5 extrinsic parameters 
2572: $\rho_{\rm max, 5}$, as described in the text. [Notice that for the spin configuration 
2573: $(60^o, 90^o, 60^o, 0^o)$ and masses $(10+10)M_\odot$ and $(15+5)M_\odot$, due to 
2574: a pole in the Pad\'e-approximant flux, we apply the Pad\'e resummation only to 
2575: the non-spinning terms in the flux.]    
2576: \label{overlapnoFL}}
2577: \end{table}
2578: 
2579: \begin{table}
2580: \begin{tabular}{c|ccccccc}
2581: \hline
2582: \hline
2583: $(\theta_{\rm S1}, \phi_{\rm S2}, \theta_{\rm S1}, \phi_{\rm S2}) $
2584: & $f_{\rm fin}^{\rm SET}$ (Hz) & 
2585: $f_{\rm LSSO}^{\rm SET}$(Hz) & 
2586: ${\rm R}_{\rm fin}^{\rm SET}/M$ & $f_{\rm fin}^{\rm SEP}$(Hz) & 
2587: $f_{\rm LSSO}^{\rm SEP}$(Hz) & 
2588: ${\rm R}_{\rm fin}^{\rm SEP}/M$ & $\rho_{\rm max, 5}$ \\
2589: \hline
2590: \hline
2591: \multicolumn{8}{l}{\hspace{6cm}$(10+10)M_\odot$}\\
2592: \hline
2593: \hline
2594: $(60^o, 90^o, 60^o, 0^o)$ & 632 & 872 & 2.6 & 616 & 1252 &2.6 & 0.8566 \\ 
2595: $(120^o, 90^o, 120^o, 0^o)$ & 185 & 185 & 6.7 & 186 & 185 &6.7 & 0.9762 \\ 
2596: \hline
2597: \hline
2598: \multicolumn{8}{l}{\hspace{6cm}$(15+5)M_\odot$} \\
2599: \hline
2600: \hline
2601: $(60^o, 90^o, 60^o, 0^o)$ & 743 & 767 & 2.3 & 661 & 772 &2.5 & 0.8232 \\ 
2602: $(120^o, 90^o, 120^o, 0^o)$ & 178 & 177 & 7.0 & 179 & 178 &6.9 & 0.9913 \\ 
2603: \hline
2604: \hline
2605: \end{tabular}
2606: \caption{Effect of Pad\'e and Taylor flux on the binary evolution and 
2607: waveforms by comparing SEHTF(3,3.5), used as target model, and 
2608: SEHPF(3,3.5), used as template model, for several binary masses and initial 
2609: spin orientations. 
2610: The two black holes are assumed to carry maximal spins $\chi_1=\chi_2=1$. 
2611: The evolution starts at $f_{\rm in} = 30$ Hz. In the first three columns 
2612: we list the ending frequency, the LSSO frequency  and the 
2613: BH radial separation at $t_{\rm fin}$ for the template model and 
2614: in the second three columns we show the same quantities but for the target 
2615: model. The last two columns contains the overlap maximized only over 2 extrinsic 
2616: parameters $\rho_{\rm max,2}$ and maximed over 5 extrinsic parameters 
2617: $\rho_{\rm max, 5}$, as described in the text. [Notice that for the spin configuration 
2618: $(60^o, 90^o, 60^o, 0^o)$ and masses $(10+10)M_\odot$ and $(15+5)M_\odot$, due to 
2619: a pole in the Pad\'e-approximant flux, we apply the Pad\'e resummation only to 
2620: the non-spinning terms in the flux.]  
2621: \label{overlapflux}}
2622: \end{table}
2623: 
2624: 
2625: \begin{table}
2626: \begin{tabular}{c|ccccccc}
2627: \hline
2628: \hline
2629: $(\theta_{\rm S1}, \phi_{\rm S2}, \theta_{\rm S1}, \phi_{\rm S2}) $
2630: & $f_{\rm fin}^{\rm SEP, noQM}$ (Hz) & $f_{\rm LSSO}^{\rm SEP, noQM}$(Hz) & 
2631: ${\rm R}_{\rm fin}^{\rm SEP, noQM}/M$ & $f_{\rm fin}^{\rm SEP}$(Hz) & 
2632: $f_{\rm LSSO}^{\rm SEP}$(Hz) & ${\rm R}_{\rm fin}^{\rm SEP}/M$ 
2633: & $\rho_{\rm max, 5}$ \\
2634: \hline
2635: \hline
2636: \multicolumn{8}{l}{\hspace{6cm}$(10+10)M_\odot$}\\
2637: \hline
2638: \hline
2639: $(60^o, 90^o, 60^o, 0^o)$ & 650 & 1257 & 2.6 & 633 & 872 & 2.6 & 0.9959 \\ 
2640: $(120^o, 90^o, 60^o, 0^o)$ & 185 & 184 & 6.8 & 186 & 186 & 6.7 & 0.9988 \\ 
2641: \hline
2642: \hline
2643: \multicolumn{8}{l}{\hspace{6cm}$(15+5)M_\odot$} \\
2644: \hline
2645: \hline
2646: $(60^o, 90^o, 60^o, 0^o)$ & 702 & 766 & 2.3 & 743 & 767 & 2.3 & 0.9823 \\ 
2647: $(120^o, 90^o, 120^o, 0^o)$ & 178 & 178 & 7.0 & 178 & 177 & 7.0 & 0.9979 \\ 
2648: \hline
2649: \hline
2650: \end{tabular}
2651: \caption{Effect of quadrupole-monopole (QM) interaction on the binary evolution and 
2652: waveforms by comparing SEHPF(3,3.5) with QM interaction, 
2653: used as target model, and SEHPF(3,3.5) without QM terms, 
2654: used as template model, for several binary masses and initial spin orientations. 
2655: The two black holes are assumed to carry maximal spins $\chi_1=\chi_2=1$. 
2656: The evolution starts at $f_{\rm in} = 30$ Hz. In the first three columns 
2657: we list the ending frequency, the LSSO frequency  and the 
2658: BH radial separation at $t_{\rm fin}$ for the template model and 
2659: in the second three columns we show the same quantities but for the target 
2660: model. The last two columns contains the overlap maximized only over 2 extrinsic 
2661: parameters $\rho_{\rm max,2}$ and maximed over 5 extrinsic parameters 
2662: $\rho_{\rm max, 5}$, as described in the text. [Notice that for the spin configuration 
2663: $(60^o, 90^o, 60^o, 0^o)$ and masses $(10+10)M_\odot$ and $(15+5)M_\odot$, due to 
2664: a pole in the Pad\'e-approximant flux, we apply the Pad\'e resummation only to 
2665: the non-spinning terms in the flux.]   
2666: \label{overlapnoQM}}
2667: \end{table}
2668: 
2669: 
2670: \begin{table}
2671: \begin{tabular}{c|ccccccc}
2672: \hline
2673: \hline
2674: $(\theta_{\rm S1}, \phi_{\rm S2}, \theta_{\rm S1}, \phi_{\rm S2}) $
2675: & $f_{\rm fin}^{\rm SEP, adiab}$ (Hz) & 
2676: $f_{\rm LSSO}^{\rm SEP, adiab}$(Hz) & 
2677: ${\rm R}_{\rm fin}^{\rm SEP, adiab}/M$ & $f_{\rm fin}^{\rm SEP}$(Hz) & 
2678: $f_{\rm LSSO}^{\rm SEP}$(Hz) & 
2679: ${\rm R}_{\rm fin}^{\rm SEP}/M$ & $\rho_{\rm max, 5}$ \\
2680: \hline
2681: \hline
2682: \multicolumn{8}{l}{\hspace{6cm}$(10+10)M_\odot$}\\
2683: \hline
2684: $(60^o, 90^o, 60^o, 0^o)$ & 636 & 1185 & 2.6 & 633 & 872 & 2.7 & 0.9666 \\ 
2685: $(120^o, 90^o, 120^o, 0^o)$ & 186 & 185 & 6.7 & 186 & 186 & 6.7 & 0.9932  \\ 
2686: \hline
2687: \hline
2688: \multicolumn{8}{l}{\hspace{6cm}$(15+5)M_\odot$} \\
2689: \hline
2690: \hline
2691: $(60^o, 90^o, 60^o, 0^o)$ & 699 & 827 & 2.3 & 743 & 767 & 2.3 & 0.9665  \\ 
2692: $(120^o, 90^o, 120^o, 0^o)$ & 177 & 177 & 7.0 & 178 & 177 & 7.0 &  0.9914  \\ 
2693: \hline
2694: \hline
2695: \end{tabular}
2696: \caption{Effect of assuming that spins evolve adiabatically. 
2697: We compare SEHPF(3,3.5), used as target model, and SEHPF(3,3.5) obtained 
2698: by averaging the spin couplings over an orbit, 
2699: as template model, for several binary masses and initial spin orientations. 
2700: The two black holes are assumed to carry maximal spins $\chi_1=\chi_2=1$. 
2701: The evolution starts at $f_{\rm in} = 30$ Hz. In the first three columns 
2702: we list the ending frequency, the LSSO frequency  and the 
2703: BH radial separation at $t_{\rm fin}$ for the template model and 
2704: in the second three columns we show the same quantities but for the target 
2705: model. The last two columns contains the overlap maximized only over 2 extrinsic 
2706: parameters $\rho_{\rm max,2}$ and maximed over 5 extrinsic parameters 
2707: $\rho_{\rm max, 5}$, as described in the text. [Notice that for the spin configuration 
2708: $(60^o, 90^o, 60^o, 0^o)$ and masses $(10+10)M_\odot$ and $(15+5)M_\odot$, due to 
2709: a pole in the Pad\'e-approximant flux, we apply the Pad\'e resummation only to 
2710: the non-spinning terms in the flux.]   
2711: \label{overlapadiab}}
2712: \end{table}
2713: 
2714: 
2715: \begin{table*}
2716: \begin{tabular}{c|c|ccc|ccc}
2717: \hline
2718: \hline
2719: $(\theta_{\rm S1}, \phi_{\rm S1}, \theta_{\rm S2}, \phi_{\rm S2}) $
2720:   & $[\delta E_H]_{f<40\,{\rm Hz}}/M$  & $f_{\rm LSSO}$ (Hz) & $[\delta E_H]_{\rm LSSO}^{40\,{\rm Hz}}/M$ 
2721:   & $\left[|\mathbf{J}|/E^2\right]_{\rm LSSO}$ 
2722:   & $f_{\rm fin}$ &
2723:    $\left[\delta E_H\right]_{\rm fin}^{40,{\rm Hz}}/M$
2724: &$\left[|\mathbf{J}|/E^2\right]_{\rm fin}$\\
2725: \hline
2726: \hline
2727: \multicolumn{8}{l}{\hspace{6cm}$(15+15)M_\odot$, 3PN}\\
2728: \hline
2729: \hline
2730: nospin &   0.0082 & 190 & 0.0107 & 0.82 & 325 & 0.0182 & 0.77  \\
2731: \hline
2732: %
2733: ($0^\circ$,$0^\circ$,$0^\circ$,$0^\circ$) &
2734: 0.0086 & (1430) &   $-$ & $-$ & 474 & 0.0527 &  0.96 \\
2735: %
2736: ($180^\circ$,$0^\circ$,$180^\circ$,$0^\circ$) &
2737: 0.0077 & 97 &   0.0033 & 0.51 & 194 & 0.0064 &  0.47 \\
2738: %
2739: ($60^\circ$,$90^\circ$,$60^\circ$,$0^\circ$) &
2740: 0.0084 & (760) &  $-$ & $-$ & 440 & 0.0352 &  0.91 \\
2741: %
2742: ($120^\circ$,$90^\circ$,$120^\circ$,$0^\circ$) &
2743: 0.0079 & 123 &   0.0054 & 0.74 & 242 & 0.0101 &  0.70 \\
2744: \hline\hline
2745: \multicolumn{8}{l}{\hspace{6cm}$(15+5)M_\odot$, 3PN}\\
2746: \hline\hline
2747: nospin & 0.0048 &  265  & 0.0084 & 0.62 & 484 & 0.0141 & 0.58 \\
2748: \hline
2749: %
2750: ($0^\circ$,$0^\circ$,$0^\circ$,$0^\circ$) &
2751: 0.0049 & (1442) &   $-$ & $-$ & 819 & 0.0493 &  0.95 \\
2752: %
2753: ($180^\circ$,$0^\circ$,$180^\circ$,$0^\circ$) &
2754: 0.0046 & 140 &   0.0034 & 0.14 & 289 & 0.0054 &  0.11 \\
2755: %
2756: ($60^\circ$,$90^\circ$,$60^\circ$,$0^\circ$) &
2757: 0.0049& (793) &  $-$ & $-$ & 719  & 0.0294 &  0.91 \\
2758: %
2759: ($120^\circ$,$90^\circ$,$120^\circ$,$0^\circ$) &
2760: 0.0047 & 
2761: 177 &   0.0049 & 0.62 & 351 & 0.0080 &  0.60 \\
2762: \hline
2763:  \end{tabular}
2764: \caption{Energy released and the magnitude of angular momentum
2765: through the evolution (with spin-spin terms ignored). For non-spinning
2766: binaries, and four configurations of maximally spinning binaries, we
2767: give the energy released below 40\,Hz, from 40\,Hz up to the LSSO, and
2768: from 40\,Hz up to to the end of the evolution. [In some cases the
2769: evolution stops before LSSO can be reached.] We also show the
2770: corresponding values of $|\mathbf{J}|/E^2$. Note that these results 
2771: do not include the ring-down phase. 
2772: \label{tab:EJ}}
2773: \end{table*}
2774: 
2775: \clearpage 
2776: 
2777: 
2778: \begin{figure}
2779: \begin{center}
2780: \epsfig{file=eomegaH.eps,width=0.95\sizetwofig,angle=0} \hspace{0.5cm}
2781: \epsfig{file=eomega.eps,width=0.95\sizetwofig,angle=0} 
2782: \caption{
2783: \label{fig:eomega}
2784: The energy for circular orbits as function of the frequency evaluated using 
2785: the PN-expanded Hamiltonian (left panel) and the 
2786: PN-expansion of the analytically computed function given by Eq.~(\ref{s1}) (right panel)  
2787: at various PN orders for maximal spins and equal mass binaries.}
2788: \end{center}
2789: \end{figure}
2790: 
2791: \vspace{0.3cm}
2792: 
2793: \begin{figure}
2794: \begin{center}
2795: \epsfig{file=eomegaEOB.eps,width=0.95\sizetwofig,angle=0}
2796: \caption{
2797: \label{fig:eomegaEOB}
2798: The energy for circular orbits as function of the frequency evaluated from the 
2799: EOB Hamiltonian at various PN orders for maximal spins and equal mass binaries.}
2800: \end{center}
2801: \end{figure}
2802: 
2803: \vspace{0.1cm}
2804: 
2805: \begin{figure}
2806: \begin{center}
2807: \epsfig{file=e-a.eps,width=0.95\sizetwofig,angle=0} \hspace{0.5cm}
2808: \epsfig{file=omega-a.eps,width=0.95\sizetwofig,angle=0}
2809: \caption{
2810: \label{fig:e-a}
2811: The energy (left panel) and the frequency (right panel) 
2812: at the LSSO as function of $\chi_L/M^2 \equiv
2813: S_{\rm eff} \cdot \hL/M^2$ in the
2814: equal mass case for EOB Hamiltonian and PN-expanded 
2815: analytically computed function ${E}(\Omega)$ [see right panel of Fig.~\ref{fig:eomega}]. 
2816: The horizontal dashed line in the right panel marks the highest LSSO angular frequency 
2817: for BBHs with total mass in the range $10\mbox{--}40 M_\odot$, assuming the 
2818: LIGO frequency band $40 \leq f_{\rm GW} \leq 240 Hz$.}
2819: \end{center}
2820: \end{figure}
2821: 
2822: \vspace{0.1cm}
2823: 
2824: \begin{figure}
2825: \begin{center}
2826: \epsfig{file=snr.eps,width=0.95\sizetwofig,angle=0} 
2827: \caption{\label{snr}
2828: Signal-to-noise ratio versus binary total mass at 100 Mpc for equal-mass binaries with LSSO 
2829: determined by the 3PN-EOB Hamiltonian.}
2830: \end{center}
2831: \end{figure}
2832: 
2833: \vspace{0.1cm}
2834: 
2835: 
2836: \begin{figure}
2837: \begin{center}
2838: \epsfig{file=CompT_ex_co_em.eps,width=0.95\sizetwofig,angle=0} \hspace{0.5cm}
2839: \epsfig{file=CompP_ex_co_em.eps,width=0.95\sizetwofig,angle=0}
2840: \caption{Newton-normalized flux in the equal-mass case with both BH spins aligned 
2841: (and maximal $\chi = \chi_1=\chi_2$) 
2842: with orbital angular momentum when T-approximants (left panel) and (upper-diagonal) 
2843: P-approximants (right panel) are used. \label{Fig6}}
2844: \end{center}
2845: \end{figure}
2846: 
2847: \vspace{0.1cm}
2848: 
2849: \begin{figure}
2850: \begin{center}
2851: \epsfig{file=CompT_ex_cr_em.eps,width=\sizetwofig,angle=0} \hspace{0.5cm}
2852: \epsfig{file=CompP_ex_cr_em.eps,width=\sizetwofig,angle=0}
2853: \caption{Newton-normalized flux in the equal-mass case with both BH spins antialigned
2854: (and maximal $\chi = \chi_1=\chi_2$) with orbital angular momentum 
2855: when T-approximants (left panel) and (upper-diagonal) P-approximants (right panel) are used. 
2856: \label{Fig7}}
2857: \end{center}
2858: \end{figure}
2859: 
2860: \vspace{0.1cm}
2861: 
2862: \begin{figure}
2863: \begin{center}
2864: \epsfig{file=Comparison_Flux3PN.eps,width=\sizetwofig,angle=0} 
2865: \caption{Comparison between the T- and (upper diagonal) P-approximant 
2866: Newton-normalized flux in the equal mass case 
2867: at 3PN and 3.5PN order.\label{Fig5}}
2868: \end{center}
2869: \end{figure}
2870: 
2871: \vspace{0.1cm}
2872: 
2873: \begin{figure}
2874: \includegraphics[width=0.9\textwidth]{rdot.eps}
2875: \caption{Oscillations in $\dot{r}$ when spin-spin interactions are
2876: present, in $(10+10)M_{\odot}$ (left panel) and $(15+5)M_{\odot}$
2877: (right panel) binaries. Dark curves show $\dot{r}/(r\omega)$ as
2878: functions of $f_{\rm GW}$ when both spin-orbit and spin-spin
2879: interactions are take into account, while light curves show the same
2880: quantity when only spin-orbit interactions are included.
2881: \label{fig:rdot}}
2882: \end{figure}
2883: 
2884: \vspace{0.1cm}
2885: 
2886: \begin{figure}
2887: \includegraphics[width=0.45\textwidth]{wf-rel-orb.eps}
2888: \caption{\label{fig:rel-orb-phase} Relative orbital phase measured with respect to $\mathbf{S}_{\rm tot}
2889: \times \hat{\mathbf{L}}_N $.  For maximally spinning $(15+15)M_\odot$
2890: binaries, we start evolution at $40\,$Hz, with
2891: $(\theta_{S_1},\phi_{S_1};\theta_{S_2},\phi_{S_2})=(60^\circ,90^\circ;60^\circ,0^\circ)$,
2892: and orbital phases $\phi_{\rm orb}=0$ and $\pi/2$, and plot the
2893: difference $\Delta\phi_{\rm orb}$. }
2894: \end{figure}
2895: 
2896: \vspace{0.1cm}
2897: 
2898: \begin{figure}
2899: \includegraphics[width=0.9\textwidth]{ss-non-adiabatic.eps}
2900: \caption{Comparison between waveforms from configurations with
2901: different initial orbital phases.  For a $(15+15)M_\odot$ maximally
2902: spinning binary, we start evolution at $40\,$Hz, with
2903: $(\theta_{S_1},\phi_{S_1};\theta_{S_2},\phi_{S_2})=(60^\circ,90^\circ;60^\circ,0^\circ)$,
2904: and orbital phases $\phi_{\rm orb}=0$ and $\pi/2$, and compare the
2905: waveforms detected with $(F_+,F_\times;\Theta,\varphi)=
2906: (1,0;\pi/4,0)$.  \label{fig:orbphase}}
2907: \end{figure}
2908: 
2909: 
2910: \vspace{0.1cm}
2911: 
2912: \begin{figure}
2913: \includegraphics[width=0.9\textwidth]{compare_energy_eob.eps}
2914: \caption{
2915: \label{fig:comp:eloss}
2916: Comparison between the three different prescriptions, $\delta E_I$
2917: (dashed curves), $\delta E_h$ (dotted curves) and $\delta E_H$ (solid
2918: curves), for calculating energy losses. We use Newtonian-order
2919: radiation reaction in the left panel, and Pad\'e at 3.5PN order in the
2920: right panel. We use the EOB Hamiltonian at 3PN order}
2921: \end{figure}
2922: 
2923: \vspace{0.1cm}
2924: 
2925: 
2926: \begin{figure}
2927: \begin{tabular}{c}
2928: \includegraphics[width=0.9\textwidth]{energy-release.eps} \\
2929: \vspace{0.2cm}\\
2930: \hspace{0.15cm}\includegraphics[width=0.9\textwidth]{angular-momentum.eps}
2931: \end{tabular}
2932: \caption{
2933: Accumulative energy release (upper panels) and instantaneous
2934: values of $|\mathbf{J}|/E^2$ (lower panels) of $(15+15)M_{\odot}$
2935: (left panels) and $(15+5)M_{\odot}$ (right panels) binaries. LSSO
2936: frequencies for the anti-aligned, generic-down, and non-spinning
2937: configurations are shown in vertical grid lines, while LSSOs of
2938: generic-up and aligned configurations are above the ranges of our
2939: plots. [Spin-spin terms are not included in these evolutions.] 
2940: We use the SEHPF(3,3.5) model. 
2941:  \label{fig:eloss}
2942: }
2943: \end{figure}
2944: 
2945: \vspace{0.1cm}
2946: 
2947: \begin{figure}
2948: \includegraphics[width=0.9\textwidth]{ending-wf.eps}
2949: \caption{
2950: \label{fig:ringdown}
2951: Inspiral waveforms (which end at $t=0$ in our plot) matched to ring-down 
2952: waveforms for non-spinning (light solid curve), and half-maximally
2953: spinning $(15+15)M_\odot$ binaries (left panel) and $(15+5)M_\odot$ binaries (right panel) 
2954: in the generic-up (dark solid curve) and generic-down (dark dashed curve) configurations. We start
2955: our evolutions at $40\,$Hz, and use $(F_+,F_\times;\Theta,\varphi)=(1,0;\pi/4,0)$. 
2956: In the plot we mark the position of the LSSO with solid curves. The waveforms have been 
2957: shifted in time such that the end of the inspiral occurs at $t/M =0$.  
2958: }
2959: \end{figure}
2960: 
2961: \end{document}
2962: 
2963: 
2964: 
2965: 
2966: 
2967: 
2968: 
2969: 
2970: 
2971: 
2972: 
2973: 
2974: