1: %\documentclass[preprint,showpacs,floatfix,aps]{revtex4}
2: \documentclass[twocolumn,showpacs,floatfix,aps]{revtex4}
3: \usepackage[dvips]{graphics}
4:
5:
6:
7:
8: \def \beq {\begin{equation}}
9: \def \eeq {\end{equation}}
10:
11: \usepackage{amssymb}
12:
13: \begin{document}
14:
15:
16: \title{Gravitational wave recoil in Robinson-Trautman spacetimes }
17:
18: \author{Rodrigo P. Macedo}
19: \email{romacedo@fma.if.usp.br}
20: \affiliation{
21: Instituto de F\'\i sica, Universidade de S\~ao Paulo, \\
22: C.P. 66318, 05315-970 S\~ao Paulo, SP, Brazil.}
23: \author{Alberto Saa}
24: \email{asaa@ime.unicamp.br}
25: \affiliation{
26: Departamento de Matem\'atica Aplicada, Universidade Estadual de Campinas, \\
27: C.P. 6065, 13083-859 Campinas, SP, Brazil.}
28:
29:
30:
31: \begin{abstract}
32: We consider the gravitational recoil due to non-reflection-symmetric gravitational wave
33: emission in the context of axisymmetric Robinson-Trautman spacetimes. We show that regular
34: initial data evolve generically into a final configuration corresponding to a
35: Schwarzschild black-hole moving with constant speed.
36: For the case of (reflection-)symmetric initial configurations, the mass of the remnant black-hole and the
37: total energy radiated away are completely determined by the initial data, allowing us to
38: obtain analytical expressions for some recent numerical results
39: that have been appeared in the literature. Moreover,
40: by using the Galerkin spectral method to analyze the non-linear regime
41: of the Robinson-Trautman equations, we show that
42: the recoil velocity can be estimated with good accuracy from
43: some asymmetry measures (namely the first odd moments) of the initial data.
44: The extension for the non-axisymmetric case and
45: the implications of our results for realistic situations involving head-on collision
46: of two black holes are also discussed.
47: \end{abstract}
48:
49: \pacs{04.30.Db, 04.25.dg, 04.70.Bw}
50: \maketitle
51: \section{Introduction}
52: The possibility that a body recoils while emitting gravitational radiation has been known
53: for decades\cite{recoil}. This problem has been considered in the literature by means of
54: many approximated and
55: semi-analytical methods as,
56: for instance, the particle approximation\cite{particle}, post-Newtonian methods\cite{postNewt},
57: and the Close-Limit Approximation\cite{closed}, leading to typical recoil velocities of
58: few hundreds of km/s for some realistic cases. Such conclusions, however, have
59: changed drastically
60: due to some recent advances in numerical relativity\cite{numrel}. In particular,
61: recent numerical simulations\cite{merger} of the merging process of binary black-holes indicate that asymmetrical gravitational wave
62: emission can indeed induce the merger remnant to recoil with velocities up to several
63: thousands of km/s.
64: The physical nature and possible implications of such considerably higher
65: gravitational recoil are now under intense investigation (see, for instance, \cite{consequences}).
66: The calculation of the recoil velocity as a function of the black-holes initial conditions is a
67: particularly important hard task. Since the full non-linear regime of Einstein equations is
68: extremely intricate and costly to analyze, some approximated or ``empirical'' formulas relating
69: the recoil velocity and the initial data have been proposed\cite{Kick}.
70:
71: The Robinson-Trautman (RT) spacetime \cite{RT} is perhaps the simplest
72: solutions of General Relativity which can be interpreted as an isolated gravitational
73: radiating system and, hence, it is certainly pertinent to the study of the gravitational
74: recoil effect. However, despite the many strong mathematical results on the RT solutions available
75: in the literature, only a few exact examples of RT spacetime are indeed known
76: in explicit form (see, for references, \cite{book}).
77: It is known, nevertheless, that a regular initial data, corresponding typically
78: to a compact body surrounded by gravitational waves, will evolve smoothly according to
79: the RT equation into a final
80: state corresponding to a remnant Schwarzschild black-hole\cite{Chrusciel}, which
81: can be at rest or moving with constant speed.
82: Our aim here is to go a step further in the characterization of such final evolution state as
83: function of the initial conditions. Our results are motivated and checked by some numerical
84: analysis.
85: The Robinson-Trautman partial differential equation
86: has been analyzed numerically in the recent literature\cite{RTnum}, being
87: particularly suitable to be numerically solved
88: by means of spectral methods\cite{spectral,rio,rio1}. We will follow Oliveira and Dami\~ao Soares\cite{rio,rio1}
89: and adopt the Galerkin method\cite{galerkin} for our analysis. However, as we will show, we will implement
90: it in a different way that will allow us to get simpler equations and a better accuracy.
91:
92: The present paper has four Sections and one Appendix.
93: In the next Section, the main aspects of axisymmetric RT spacetimes are presented briefly. We show,
94: in particular,
95: how to read from the final state of the RT evolution the mass and speed of the remnant
96: black-hole. It is also shown that, as expected,
97: for (reflection-)symmetric initial data there is no radiation recoil.
98: In such a case,
99: the mass of the remnant black-hole and the
100: total energy radiated away are completely determined by the initial data, allowing us to
101: establish analytical expressions for the results about the total radiated energy
102: obtained numerically in \cite{rio} and \cite{rio1}.
103: Section III is devoted to the study of generic axisymmetric initial data. We show that a typical RT
104: evolution can lead to a gravitational recoil. A Galerkin projection method is used to calculate the
105: final black-hole speed. We show also how the final recoil velocity can be estimated with good
106: accuracy from some
107: asymmetry measures of the initial configuration, namely the first odd moments of the initial data.
108: In the
109: last Section, we discuss the
110: physical interpretation of the typical initial data considered in this work, emphasizing
111: their relation with the problem of frontal collision of two black-holes. The extension of our results
112: to the non-axisymmetric case is also commented in the last section.
113: The Appendix presents a direct proof of a mathematical result used in Section II,
114: namely that, for regular
115: initial data, the final state of the
116: RT evolution does correspond generically to a Schwarzschild black hole moving with constant speed.
117:
118:
119: \section{Axisymmetric RT spacetime}
120: The standard form of the
121: Robinson-Trautman (RT) metric in the usual spherical radiation coordinates $(u,r,\theta,\phi)$ reads\cite{book}
122: \beq
123: \label{RT}
124: ds^2 = -\left(K - 2\frac{m_0}{r} - r (\ln Q^2)_u\right)du^2 - 2dudr +\frac{r^2}{Q^2}d\Omega^2,
125: \eeq
126: where $Q=Q(u,\theta,\phi)$, $m_0$ is a constant mass parameter, and
127: $d\Omega^2$ and $K$ stand for, respectively, the metric of the unit sphere and
128: the gaussian curvature of the surface corresponding to $r=1$ and $u=u_0$ constant, which is given by
129: \beq
130: K = Q^2\left(1 + \frac{1}{2}\nabla^2_\Omega \ln Q^2 \right),
131: \eeq
132: with $\nabla^2_\Omega$ corresponding to the Laplacian on the unit sphere.
133: Vacuum Einstein's equations for the metric (\ref{RT}) implies the Robinson-Trautman non-linear
134: partial differential equation\cite{book}
135: \beq
136: 6m_0\frac{\partial}{\partial u} \left( \frac{1}{Q^2}\right) = \nabla^2_\Omega K.
137: \eeq
138: In this paper,
139: we will focus on axisymmetric spacetimes and hence we will assume hereafter that $Q=Q(u,\theta)$. By
140: introducing $x=\cos\theta$ one has
141: \beq
142: \label{lambda}
143: K = Q^2 + Q \frac{\partial }{\partial x} \left[(1-x^2) Q_x \right] - (1-x^2)Q_x^2
144: \eeq
145: and
146: \beq
147: \label{RT1}
148: 6m_0\frac{\partial}{\partial u} \left( \frac{1}{Q^2}\right) = \left[(1-x^2)K_x \right]_x,
149: \eeq
150: where
151: \begin{eqnarray}
152: \label{lambdax}
153: \left[(1-x^2)K_x \right]_x &=& (1-x^2)^2\left( QQ_{xxxx} - Q_{xx}^2 \right) \\
154: &-& 8(x-x^3)QQ_{xxx}
155: - 4(1-3x^2)QQ_{xx}. \nonumber
156: \end{eqnarray}
157: Integrating (\ref{RT1}) and assuming a regular gaussian curvature $K$
158: one has
159: \beq
160: \label{constr}
161: \frac{d}{d u}\int_{-1}^1 \frac{dx}{Q^2(u,x)} = 0,
162: \eeq
163: implying that the quantity $q_0 = \int_{-1}^1 Q^{-2} dx $ is constant
164: along the solutions of (\ref{RT1}). Notice that, from (\ref{RT}), the regularity of the surface
165: $u$ and $r$ constants precludes us of having $Q=0$.
166: The regularity of the gaussian curvature $K$, on the other hand,
167: requires $0<Q<\infty$.
168: We normalize our data in order to have $q_0=2$,
169: implying that the area of the surface corresponding to $r$ and $u$ constants is always
170: $4\pi r^2$ along the
171: $u$-evolution governed by (\ref{RT1}).
172:
173: Several classical results assure that, given a geometrically regular initial
174: data $Q(0,x)$, the solution of (\ref{RT1}) approaches asymptotically a stationary ($Q_u=0$) regime.
175: The stationary solutions of (\ref{RT1}) are such that
176: \beq
177: (1-x^2)K_x = A = {\rm\ constant},
178: \eeq
179: leading to
180: \beq
181: K = A\,{\rm arctanh}\, x + B,
182: \eeq
183: where $B$ is another constant. Regularity of $K$ on the interval $[-1,1]$
184: requires necessarily $A=0$. On the other hand,
185: Eq. (\ref{lambda}) implies that the regular $Q$ solutions for which $K$ is constant
186: are such that $Q_{xx}=0$ (see the Appendix for a direct proof). Therefore,
187: the stationary solutions of (\ref{RT1}) are always of the form
188: $Q = a + bx$ with $a$ and $b$ constants. Nevertheless, our choice of $q_0=2$ yields
189: $a^2-b^2=1$. We choose in this work
190: a parametrization such that $a=\cosh \alpha$ and $b=\sinh \alpha$.
191:
192:
193: Given a normalized regular
194: initial data $Q(0,x)$, the asymptotic solution of (\ref{RT1}) will be always of the form
195: $Q(\infty,x)= \cosh\alpha + x\sinh\alpha$.
196: The final configuration is, hence, completely characterized by the sole parameter $\alpha$.
197: In order to unveil its physical role,
198: let us consider the Bondi's mass function\cite{mass}
199: \beq
200: \label{mass}
201: M(u) = \frac{m_0}{2}\int_{-1}^{1} \frac{dx}{Q^3(u,x)}
202: \eeq
203: which
204: has several desirable properties to define an ``instantaneous'' mass for the
205: solutions of (\ref{RT1}), see, for instance, \cite{rio}.
206: In particular, we have that $M(u) \ge m_0$ for normalized initial data and, for $u\rightarrow\infty$, it reduces to
207: \begin{eqnarray}
208: M(\infty) &=& \frac{m_0}{2}\int_{-1}^1 \frac{dx}{(\cosh\alpha + x\sinh\alpha )^3} = m_0\cosh\alpha \nonumber \\
209: && \quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad = \frac{m_0}{\sqrt{1-v^2}},
210: \end{eqnarray}
211: where $v=\tanh\alpha$ can be interpreted as the final velocity along the $z$
212: axis of the remnant
213: black-hole\cite{mass}.
214:
215: The Bondi's mass (\ref{mass}) corresponds to the temporal component of the Bondi's
216: four-momentum, which for generic (non-axisymmetric) RT solutions is given by\cite{mass}
217: \beq
218: P_a(u) = \frac{m_0}{4\pi} \int_{S^2} \frac{\eta^a}{ Q^3(u,\theta,\phi)} dS,
219: \eeq
220: where $S^2$ is the unit sphere spanned by the usual coordinates $\theta$ and $\phi$
221: and with area element $dS$, and $a=0,1,2,3$, with $\eta^0=1$ and $\eta^i$ being the
222: radial
223: three-vector directed to the point $(\theta,\phi)$ on the unit sphere. For axisymmetric configurations,
224: the non-vanishing components of the Bondi's four-momentum are $P_0(u) = M(u)$ and
225: \beq
226: P_3(u) = \frac{m_0}{2}\int_{-1}^{1} \frac{ x}{Q^3(u,x)} dx,
227: \eeq
228: which corresponds to the momentum carried by the solution along the $z$ axis.
229: For normalized initial data one has for $u\rightarrow\infty$
230: \begin{eqnarray}
231: P_a(\infty) = \frac{m_0} {\sqrt{1-v^2}}(1,0,0,-v),
232: \end{eqnarray}
233: reinforcing the interpretation of $v$ as the final velocity of the remnant
234: black-hole.
235:
236:
237: Notice that, for symmetric (even) initial data $Q(0,x)$, Eq. (\ref{lambdax}) implies that the solutions
238: $Q(u,x)$
239: of (\ref{RT1}) are necessarily even for $u\ge 0$, establishing that there is no
240: gravitational recoil ($v=0$) in this case. Such a behavior is, of course, in full agreement with the expectation that gravitation
241: recoil should be due to non-reflection-symmetric
242: gravitational wave emission. Therefore, for even
243: situations, the
244: constraint (\ref{constr}) determines completely the final evolution state.
245:
246: \subsection{Radiated energy: reflection-symmetric case}
247:
248: The fraction of the initial mass $M(0)$ radiated away along the $u$-evolution governed by
249: (\ref{RT1}) can be calculated exactly for even configurations. Following \cite{rio}, we define
250: \beq
251: \Delta = \frac{M(0)-M(\infty)}{M(0)},
252: \eeq
253: which clearly corresponds to the fraction of the initial mass lost due to gravitational wave
254: emission. For even configurations, $v=0$ and we have simply
255: \beq
256: \Delta = 1 - 2\left( \int_{-1}^{1} \frac{dx}{Q^3(0,x)}\right)^{-1}.
257: \eeq
258: It can be shown that $0\le\Delta <1$.
259: As an explicit example of this exactly soluble case,
260: let us consider the first even initial data considered in the papers
261: \cite{rio,rio1}, namely the
262: prolate spheroid corresponding to
263: \beq
264: \label{prola}
265: Q^2(0,x) = Q_0^2\left( 1-\epsilon^2x^2\right),
266: \eeq
267: with $0\le \epsilon < 1$. The constraint $q_0=2$ implies that
268: \beq
269: Q_0^2 = \frac{1}{2\epsilon}\ln \left(\frac{1+\epsilon}{1-\epsilon}\right),
270: \eeq
271: leading finally to
272: \beq
273: \label{delta}
274: \Delta = 1 - \sqrt{\frac{1-\epsilon^2}{8\epsilon^3}\ln^3 \left(\frac{1+\epsilon}{1-\epsilon}\right) } .
275: \eeq
276: This is the exact analytical expression for the curves obtained in \cite{rio} and \cite{rio1}
277: from numerical simulations.
278: For sake of
279: comparison with the results of \cite{rio,rio1}, Fig. \ref{fig1}
280: depicts a semi-log plot of $\Delta$ as
281: a function of $y=1-\epsilon$, following their conventions. A very good
282: agreement is found.
283: One can proceed in an analogous way for any other even (reflection-symmetric)
284: configuration, we will return to this issue in the last Section. The exact expression
285: for $\Delta$ is certainly valuable to the investigation of statistical properties of the
286: non-linear gravitational wave emission as those ones considered in \cite{rio,rio1}. For instance,
287: it is clear from (\ref{delta}) that
288: the non-extensive
289: \begin{figure}[h!]
290: \resizebox{1\linewidth}{!}{\rotatebox{0}{\includegraphics*{fig1.eps}}}
291: \caption{ The
292: fraction $\Delta$ of the initial Bondi's mass lost due to gravitational wave
293: emission for the initial configuration (\ref{prola}), as a function of $y=1-\epsilon$.
294: The curve is in very good agreement
295: with that one inferred from numerical results in \cite{rio,rio1}. Notice, however, that
296: the non-extensive distribution function proposed in \cite{rio,rio1} is merely an
297: approximation for $y\approx 1$, see (\ref{decay}).}
298: \label{fig1}
299: \end{figure}
300: distribution function proposed in \cite{rio,rio1} is only an
301: approximation valid for small $\epsilon$.
302: In fact, we have
303: \beq
304: \label{decay}
305: \Delta = \frac{1}{30}(1-y)^4 + \frac{32}{945}(1-y)^6 + O((1-y)^8),
306: \eeq
307: for $y\approx 1$ (or $\epsilon\approx 0$). Notice that $\Delta\rightarrow 1$ for
308: $\epsilon\rightarrow 1$.
309:
310:
311:
312:
313:
314:
315: \section{General Solutions}
316:
317: The evolution of generic initial data $Q(0,x)$ is a greater challenge. Since the gravitational
318: recoil is clearly related to the odd part of the function $Q(u,x)$,
319: one might consider in first place some asymmetry measures of
320: the initial data. The simplest ones correspond to their first odd $n$ moments
321: \beq
322: \label{qn0}
323: q_n(u) = \int_{-1}^{1} \frac{x^n}{Q^2(u,x)} dx,
324: \eeq
325: which obey $-q_0 \le q_n \le q_0$.
326: For the generic final evolution state $Q(\infty,x)= \cosh\alpha + x\sinh\alpha$, we have
327: \begin{widetext}
328: \beq
329: \label{qn}
330: q_n(\infty) = (1-v^2)\int^{1}_{-1} \frac{x^n}{(1 + vx)^2}dx =
331: - \frac{1}{v^n}\left(2-n\frac{1-v^2}{v}\ln\frac{1+v}{1-v} \right)
332: - \frac{1-v^2}{v^{n+1}}\sum_{k=2}^n\sum_{{\rm odd\,}j}^{k-1}\frac{(-1)^k}{k-1}{n \choose k} {k-1 \choose j}v^j,
333: \eeq
334: \end{widetext}
335: valid for odd $n$.
336: Fig. \ref{fig2}
337: \begin{figure}[ht]
338: \resizebox{1\linewidth}{!}{\rotatebox{0}{\includegraphics*{fig2.eps}}}
339: \caption{ Final odd moments $q_n(\infty)$, $n=1, 3, 5$, as functions of the recoil velocity
340: $v =\tanh\alpha$, as
341: given by (\ref{qn}). Notice that, for a given $0 < |v| < 1$, one has $|q_1|>|q_3|>|q_5|>\cdots$}
342: \label{fig2}
343: \end{figure}
344: shows the first final odd moments $q_n(\infty)$ as functions of the recoil velocity $v$.
345: As we will show, the relevance of the first odd moments (\ref{qn0}) rests on the fact that
346: one can construct, as a linear combination of them, a second approximately conserved quantity
347: along the solutions of the RT equation (\ref{RT1}) in the framework of the Galerkin
348: approximation.
349:
350:
351:
352:
353:
354: \subsection{The Galerkin method}
355: We introduce now a Galerkin decomposition for $Q(u,x)$
356: \beq
357: \label{gal}
358: Q(u,x) = \sum_{\ell=0}^N b_\ell(u)P_\ell(x),
359: \eeq
360: where $P_\ell(x)$ stands for the Legendre polynomials. By using standard projection
361: techniques\cite{galerkin}, Eq. (\ref{RT1}) can be written as the system of ordinary differential equations
362: \beq
363: \label{RTG}
364: \dot{b}_\ell = -\frac{2\ell+1}{24m_0}\langle Q^3\left[(1-x^2)K_x \right]_x,P_\ell \rangle,\quad \ell=0,1,\dots,N,
365: \eeq
366: where the inner product is given by
367: $\langle f,g \rangle = \int_{-1}^1 fg \,dx.$ From (\ref{lambdax}) and (\ref{gal}),
368: one can see that the functions involved in the inner product in
369: the right-handed side of (\ref{RTG}) are simple polynomials in $x$.
370: The integration can
371: be performed exactly for arbitrary $N$ (with the help of algebraic
372: manipulation software as Maple, for instance),
373: yielding $5^{\rm th}$ order polynomials
374: on the mode functions $b_\ell$. Notice that here, in contrast to the approach adopted in
375: \cite{rio,rio1}, no transcendental
376: function is involved in the Galerkin approximation.
377: Now, the Cauchy problem for the RT equation
378: corresponds basically in choosing the initial value of
379: the mode functions $b_\ell(u)$ according to
380: \beq
381: \label{DATA}
382: b_\ell(0) = \frac{2\ell + 1}{2 }\langle Q(0,x), P_\ell \rangle,
383: \eeq
384: and then to solve the Initial Value Problem (IVP) given by (\ref{RTG}).
385:
386: Equation (\ref{RTG}) has some useful properties that are independent of $N$.
387: For instance, their stationary solutions ($\dot{b}_\ell=0$) have necessarily
388: $b_\ell = 0$ for $\ell > 1$ and arbitrary (constants)
389: $b_0$ and $b_1$.
390: Indeed, for any regular initial data, the systems evolves into
391: the final state $Q(\infty,x)= b_0(\infty) P_0(x) + b_1(\infty) P_1(x)$, with
392: $b_0(\infty)^2 - b_1(\infty)^2 = 1$ for the normalized case,
393: as expected. The recoil velocity will be given simply by $v=b_1(\infty)/b_0(\infty)$.
394: Another useful property is that for an even initial data, one has
395: $b_\ell(u) =0$ for odd $\ell$
396: and, consequently, $v=0$.
397: The accuracy of the Galerkin decomposition is determined by the truncation order $N$ in
398: (\ref{gal}). It can be controlled effectively here by checking
399: the conserved quantity $q_0$ along the $u$-evolution.
400: Typically, the expansion with $N$ Legendre polynomials in (\ref{gal}) is accurate provided that
401: $\max |b_N(u)|$ be small enough.
402:
403: Finally,
404: we are able now to
405: consider the evolution of generic initial data. The recoil velocity $v$ can
406: be calculated by solving the IVP corresponding to
407: the system of ordinary differential equations (\ref{RTG}) with initial conditions (\ref{DATA}).
408: The recoil velocity determines completely the final state for normalized initial data,
409: allowing the study of any other
410: relevant quantity as, for instance, the fraction $\Delta$
411: of the initial mass radiated away as a function
412: of the non-reflection-symmetric initial data,
413: \beq
414: \label{delta11}
415: \Delta = 1 - \frac{2}{\sqrt{1-v^2}}\left( \int_{-1}^{1} \frac{dx}{Q^3(0,x)}\right)^{-1} .
416: \eeq
417: We have performed an exhaustive numerical analysis of the system
418: (\ref{RTG}). The considered initial data include the following simple but
419: representative family
420: \beq
421: \label{family}
422: Q(0,x) = Q_0 \left( 1 + \alpha x + \beta x^2 + \gamma x^3\right),
423: \eeq
424: where the
425: constant $Q_0$ is always chosen in order to ensure the normalization $q_0=2$.
426: Some particular elements of this family
427: \begin{figure*}[ht]
428: \resizebox{1\linewidth}{!}{ \includegraphics*{fig3.eps}}
429: \caption{Polar plot of some typical non-reflection-symmetric initial data $Q(0,x)$ of the family (\ref{family}). The initial condition $a$, $b$, and $c$ correspond, respectively, to the parameters $\alpha=1/2$, $\beta=1$, $\gamma=0$;
430: $\alpha= \beta=0$, $\gamma=-2/3$; and $\alpha=0 $, $\beta=4$, $\gamma=3$. The dashed lines correspond to
431: the associated gravitational radiation content (without scale), see Sect. IV.
432: }
433: \label{figcond}
434: \end{figure*}
435: are presented in Fig. \ref{figcond}.
436:
437:
438:
439: Fig. \ref{figcalc} depicts
440: \begin{figure}[ht]
441: \resizebox{1\linewidth}{!}{\rotatebox{0}{\includegraphics*{fig4.eps}}}
442: \caption{ Evolution of the modes $b_\ell(u)$ governed by (\ref{RTG}) for the
443: case $(a)$ of Fig. \ref{figcond}. $N=8$ was used, leading to an accuracy
444: (controlled by the constant $q_0=2$) of $10^{-4}$. The final evolution state
445: has $b_0=1.0197$ and $b_1=0.20017$ and, consequently, the recoil velocity is
446: $v=0.19628$ and the radiated energy fraction $\Delta=0.05420$, calculated
447: according to (\ref{delta11}).}
448: \label{figcalc}
449: \end{figure}
450: a typical evolution for the modes $b_\ell(u)$ governed by (\ref{RTG}) for
451: a particular case of the family (\ref{family}). The recoil velocity $v$ can be read from the
452: final state of the evolution for any initial data.
453: We notice that, for the family of initial conditions (\ref{family}), we always have $b_\ell(0)=0$
454: for $\ell>3$ and, in this case,
455: $N=8$ is sufficient to assure typically an accuracy (controlled by the constant $q_0(u)=2$) of the Galerkin
456: approximation up to 1\%. Initial data with high $q_n(0)$ typically require higher $N$ in order
457: to attain a given accuracy. The radiation content of the initial data can also give some clues about
458: the minimal necessary value of $N$, see Section IV.
459:
460: \subsection{Estimation of the recoil velocity}
461:
462: Despite that the IVP associated to the equation (\ref{RTG}) can be solved with
463: quite modest computational resources, an analytical estimation of the recoil velocity $v$
464: from the initial data would be certainly valuable. Since the final state of the RT
465: evolution is completely characterized by the sole parameter $v$ for normalized initial
466: data, a second conserved quantity
467: besides $q_0$ would suffice to determine completely the final state and, consequently, to
468: determine the recoil velocity $v$. Unfortunately, the RT
469: equation (\ref{RT1}) does not seem to have any other conserved quantity rather than $q_0$.
470: On the other hand, its Galerkin approximation (\ref{RTG}) does indeed have a second conserved quantity.
471: Such a new conserved quantity, however, will be only approximately constant along the solutions
472: of the full RT equation. Nevertheless, the approximation will be as good as the
473: Galerkin approximation is accurate.
474: In order to construct an explicit expression for the
475: new constant, we remind that (\ref{RT1}) implies that
476: the moments (\ref{qn0}) obey the equation
477: \beq
478: \label{dqn0}
479: 6m_0\dot{q}_n (u) = \langle x^{n},\left[(1-x^2)K_x \right]_x \rangle.
480: \eeq
481: From (\ref{lambdax}) and (\ref{gal}), we see that
482: \beq
483: \label{sum}
484: \left[(1-x^2)K_x \right]_x = \sum_{\ell=0}^{2N} a_\ell(u)x^\ell,
485: \eeq
486: where $a_\ell(u)$ are quadratic functions of the modes $b_\ell(u)$. For odd $n$,
487: the inner product in (\ref{dqn0}) will
488: select only the odd-$\ell$ terms in $x$ in the summation (\ref{sum}), leading to the following linear relation
489: between $\dot{q}_n (u)$ and $a_\ell(u)$
490: \beq
491: \label{dqn01}
492: 3m_0\dot{q}_n (u) = \sum_{{\rm odd\,}\ell}^{2N} \frac{a_{\ell}(u)}{\ell+n+1}.
493: \eeq
494: The right-handed side of (\ref{dqn01}) has exactly $N$ terms, implying, therefore, that one
495: can have at most $N$ linear independent equations of the type (\ref{dqn0}).
496: The linear relation between $\dot{q}_n$ and $a_\ell(u)$ given by (\ref{dqn01}) involves a Hilbert-type matrix\cite{hilbert} and,
497: in particular,
498: it is always possible to find $N+1$ rational numbers $\alpha_\ell$ such that
499: \beq
500: \label{alpha}
501: \frac{d}{du} \left( \sum_{\ell=1}^{N+1} \alpha_\ell q_{2\ell-1}(u) \right)= 0.
502: \eeq
503: The quantity between parenthesis is conserved along the solutions of (\ref{RTG}) and,
504: therefore, it corresponds to our second conserved quantity.
505: One could also truncate the summation in (\ref{dqn01}) in a given $\ell$, obtaining partial
506: linear combination of the odd moments that are constant along the solutions of (\ref{RTG}) up
507: to deviations proportional to max $|a_{\ell+2}(u)|$. The first of such partial linear combinations are
508: \begin{eqnarray}
509: (\ell = 0) && q_1, \\
510: (\ell = 1) && q_1 - \frac{5}{3}q_3, \\
511: \label{cond1}
512: (\ell = 3) && q_1 -\frac{14}{3}q_3 + \frac{21}{5}q_5, \\
513: && \vdots \nonumber
514: \end{eqnarray}
515: The coefficients in the above expressions and the $\alpha_\ell$ of (\ref{alpha}) can be calculated
516: in a straightforward way by using, for instance, Gauss elimination in (\ref{dqn01}). However,
517: our numerical calculations show that,
518: for the typical initial data considered here, the first odd moment $q_1$ {\em dominates}
519: over the other ones, implying that the typical variations $(q_1(0)-q_1(\infty))/q_1(0)$
520: are rather small. We notice also that the typical initial data of the family (\ref{family}) considered
521: here has $|q_1(0)| > |q_3(0)| > |q_5(0)| > \cdots$, in agreement with the magnitude of the
522: odd moments for the final state, see Fig. \ref{fig2}. This situation can fail for some
523: very specific initial conditions. For instance, if one has $|q_3(0)| > |q_1(0)| > |q_5(0)| > \cdots$,
524: $q_1$ will vary considerably along the solutions of (\ref{RTG}), but the combination
525: given by (\ref{cond1}) will be approximately constant, and so on.
526: Fig. \ref{figqn} presents
527: \begin{figure}[ht]
528: \resizebox{1\linewidth}{!}{\rotatebox{0}{\includegraphics*{fig5.eps}}}
529: \caption{ Plot of $v\times q_1(0)$ for some typical regular initial conditions.
530: The dotted line is the curve predicted by (\ref{vfinal}). The detail depicts
531: the plot of $q_1(0)\times q_1(\infty)$.
532: The assumption of $q_1(\infty)=q_1(0)$ is, typically,
533: a good approximation when $q_1(u)$ is the dominant moment.}
534: \label{figqn}
535: \end{figure}
536: numerical evidences confirming these results.
537: For practical purposes, whenever $|q_1(0)| > |q_3(0)| > |q_5(0)| > \cdots$,
538: one can assume that $q_1(\infty)\approx q_1(0)$ and estimate the final
539: recoil velocity $v$ as
540: \beq
541: \label{vfinal}
542: \frac{1}{v }\left(2- \frac{1-v^2}{v}\ln\frac{1+v}{1-v} \right)\approx - \int_{-1}^{1} \frac{x}{Q^2(0,x)}dx.
543: \eeq
544: In particular, $v$ has the opposite sign of $q_1(0)$, see Fig. \ref{figqn}.
545: We emphasize, nevertheless, that (\ref{vfinal}) will be accurate solely in the
546: cases where $q_1(u)$ is actually the dominant moment.
547:
548:
549: \section{Discussion}
550:
551: The physical properties of the initial conditions corresponding to the family (\ref{family}) can be
552: investigated by considering
553: their radiation content, which is determined by the $(1/r)$-decaying part
554: of the Riemann tensor and is proportional to the quantity\cite{book, rio2}
555: \beq
556: D(u,x) = - (1-x^2) Q^2 \partial_u \left( \frac{Q_{xx}}{Q} \right).
557: \eeq
558: With the help of (\ref{RT1}) and (\ref{lambdax}), one can show that
559: for polynomials $Q(u,x)$ in $x$,
560: the function $D(u,x)$ will be also polynomial in $x$. Moreover, $D(u,x)$
561: is an even (reflection-symmetric) function for even $Q(u,x)$.
562: The dashed lines in
563: Fig. 3 are polar plots without scale of $|D(0,x)|$ corresponding
564: to the radiation content of the associated initial data. The asymmetry in the
565: gravitational radiation emission responsible for the final recoil is clear.
566: We notice that initial data with larger $\max |D(0,x)|$ will typically require
567: a larger value of the truncation order
568: $N$ to attain a given accuracy in the Galerkin approximation. For instance,
569: case (c) of Fig. 3 requires a truncation order larger than cases (b) and (a)
570: to keep the same accuracy.
571:
572: Some cases of the family (\ref{family}) are specially
573: interesting since they are good approximations for the Brill-Lindquist
574: initial data\cite{Brill}
575: \beq
576: \label{2bh}
577: Q(0,x) = Q_0\left( \frac{1}{\sqrt{1 -wx}} + \frac{\mu}{\sqrt{1 +wx}} \right)^{-2},
578: \eeq
579: which can be interpreted as the final stage (after the horizon merging)
580: of a frontal collision of two black holes\cite{headone,headon}, with the parameters
581: $\mu\ge 0$ and $0\le w<1$ related, respectively, to the mass ratio and to the infalling relative velocity
582: of the two black-holes. The constant $Q_0$ must be chosen in order to assure $q_0=2$. We have
583: \beq
584: \label{headQ}
585: Q_0^2(\mu,w) = \frac{1+\mu^4}{1-w^2} + \frac{4\mu(1 + \mu^2)}{\sqrt{1-w^2}} + 3\frac{\mu^2}{w}
586: \ln \left(\frac{1+w}{1-w} \right).
587: \eeq
588: For $\mu=1$ (the equal masses case), the function (\ref{2bh}) is reflection-symmetric and, in this case,
589: the final state of the evolution is completely determined by the constraint $q_0=2$.
590: For $\mu \ne 1$, one can estimate the recoil velocity for this head-on collision approximation
591: by using (\ref{vfinal}). For the the initial data (\ref{2bh}) we have
592: \begin{widetext}
593: \beq
594: \label{vxw}
595: \frac{1}{v }\left(2- \frac{1-v^2}{v}\ln\frac{1+v}{1-v} \right) \approx -q_1(0)=
596: \frac{\left( \mu^{-2}-\mu^2 \right)\left(
597: \frac{1}{w^2}\ln\frac{1+w}{1-w} - \frac{2}{w-w^3}
598: \right)
599: + 8\left( \mu^{-1} -\mu\right) \left(
600: \frac{\arcsin w}{w^2} - \frac{1}{w\sqrt{1-w^2}}
601: \right)
602: }{\frac{\mu^2 + \mu^{-2}}{1-w^2} + \frac{4(\mu + \mu^{-1})}{\sqrt{1-w^2}} + \frac{3}{w}
603: \ln \left(\frac{1+w}{1-w} \right)}.
604: \eeq
605: \end{widetext}
606: For small values of $w$, the condition (\ref{vxw}) reduces to
607: \beq
608: v = \frac{\mu-1}{\mu+1}w.
609: \eeq
610: Fig. 6 shows the dependence of $v$ with $w$ for some values of $\mu$
611: \begin{figure}[ht]
612: \resizebox{0.95\linewidth}{!}{\rotatebox{0}{\includegraphics*{fig6.eps}}}
613: \caption{ Dependence of the recoil velocity $v$ with the infalling velocity
614: $w$ of the two black holes with different masses, as predicted by (\ref{vxw}), and
615: some results from numerical calculations.
616: Notice that $v\rightarrow -v$ if $\mu \rightarrow 1/\mu$.
617: }
618: \end{figure}
619: as predicted by (\ref{vxw}) and some numerical results. A very good agreement is
620: found again. It is interesting to notice that
621: \beq
622: \lim_{w\rightarrow 1} q_1(0) = 2\frac{\mu^4-1}{\mu^4+1}
623: \eeq
624: for the initial data (\ref{2bh}), implying from (\ref{vxw}) that there exists
625: a maximum recoil velocity for this configuration
626: \beq
627: \label{maxvel}
628: \lim_{w\rightarrow 1} |v| = v_{\rm max} < 1.
629: \eeq
630: In fact, Eq. (\ref{vxw}) implies that $v<w$ for any $\mu$ (see Fig. 6), a behavior
631: already noticed in the numerical analysis of \cite{headon}.
632: One can also calculate the
633: fraction (\ref{delta11}) of the initial mass radiated away for this case
634: \beq
635: \label{delta2}
636: \Delta = 1 - \frac{2}{\sqrt{1-v^2}}\frac{Q_0^3(u,w)}{h(u,w)
637: },
638: \eeq
639: where $v$ is given by (\ref{vxw}) and
640: \begin{widetext}
641: \beq
642: h(u,w) =
643: \frac{2(1+\mu^6)}{(1-w^2)^2} + \frac{8(\mu+\mu^5)}{(1-w^2)^\frac{3}{2}} + \frac{15(u^2+u^4)}{1-w^2} +
644: \frac{4(\mu+\mu^5)+40\mu^3}{\sqrt{1-w^2} } + \frac{15(\mu^2+\mu^4)}{2w}\ln\left(\frac{1+w}{1-w} \right).
645: \eeq
646: \end{widetext}
647: The aspect of the curves (\ref{delta2}) are similar to that one depicted in
648: Fig. 1.
649: In
650: particular, for
651: small $w$, one has
652: \beq
653: \Delta = \frac{3}{5}\frac{\mu(5\mu^2-8\mu+5)}{ (\mu+1)^4}w^4 + O(w^6),
654: \eeq
655: compare with (\ref{decay}). Due to (\ref{maxvel}), one has $\Delta\rightarrow 0$ irrespective of
656: $\mu$ for $w\rightarrow 1$.
657:
658:
659:
660: We finish by commenting that the
661: non-axisymmetric case $Q=Q(u,\theta,\phi)$ can also be investigated by means of a Galerkin method.
662: For such a case, the Galerkin decomposition (\ref{gal}) is based on the spherical harmonics
663: \beq
664: Q(u,\theta,\phi) = \sum_{\ell=0}^N \sum_{m=-\ell}^{\ell} b_{\ell m}(u)Y_\ell^m(\theta,\phi),
665: \eeq
666: and a system of equations equivalent to (\ref{RTG}) can be obtained. In this case,
667: the stationary regime corresponds
668: also to the case for that $b_{\ell m} = 0$ for $\ell >1$. The constant-$K$
669: final state will have the form
670: \beq
671: Q(\infty,\theta,\phi) = b_{00} + b_{10}\cos\theta + a\sin\theta\cos\phi + c\sin\theta\sin\phi,
672: \eeq
673: where $b_{00}^2 - (b_{10}^2 + a^2 + c^2)=1$
674: for the normalized case.
675: The non-vanishing coefficients now can determine the modulus and the direction of the
676: Bondi's four-momentum and, consequently, the recoil velocity of the
677: remnant. These topics are now under investigation.
678:
679: \acknowledgements
680: The authors are grateful to I.D. Soares, H. Oliveira, and R. Mosna
681: for enlightening discussions.
682: This work was supported by FAPESP and CNPq.
683:
684: \appendix
685:
686:
687:
688: \section{}
689:
690: One can check easily by a direct substitution that $Q(x) = a + bx$, with $a^2-b^2=K$, is a regular
691: solution of (\ref{lambda}).
692: Nevertheless, a stronger result holds in this case: {\em all} regular solutions of (\ref{lambda})
693: with constant $K$
694: are necessarily
695: of this form.
696: We are interested in the geometrically regular solutions ($0<Q(x)<\infty$ and
697: $|Q_x(x)|<\infty$
698: for $-1\le x\le 1$).
699: The relevant phase space is three-dimensional and spanned by $(x,Q,Q_x)$. Notice that any
700: solution such that $Q_{xx}=0$ must be constrained on the surface ${\cal L}$ of the phase space corresponding
701: to the points such that
702: \beq
703: L(x,Q,Q_x) = Q^2 - 2xQQ_x - (1-x^2)Q_x^2 - K = 0.
704: \eeq
705: One can show that, along any solution of (\ref{lambda}), one has
706: \beq
707: \label{eqa1}
708: (1-x^2)Q\frac{dL}{dx} = 2(xQ + (1-x^2)Q_x)L,
709: \eeq
710: confirming that ${\cal L}$
711: is indeed an invariant surface of (\ref{lambda}). The linear equation (\ref{eqa1}) has the
712: solution
713: \beq
714: L( x,Q(x),Q_x(x)) = A \frac{Q^2(x)}{1-x^2},
715: \eeq
716: where $A$ is a constant, implying that any solution of (\ref{lambda}) such that $L\ne 0$
717: cannot be regular in $x=\pm 1$
718:
719:
720:
721: \begin{references}
722:
723: \bibitem{recoil}W. B. Bonnor and M. A. Rotenberg, Proc. R. Soc. Lond. A. {\bf 265}, 109 (1961);
724: A. Peres, Phys. Rev. {\bf 128}, 2471 (1962); J.D. Bekenstein, Astrophys. J. {\bf 183}, 657 (1973);
725:
726: \bibitem{particle}M.J. Fitchett and S. Detweiler, MNRAS {\bf 211}, 933 (1984);
727: T. Nakamura and M. P. Haugan, Astrophys. J. {\bf 269}, 292
728: (1983); M. Favata, S. A. Hughes, and D. E. Holz, Astrophys. J.
729: {\bf 607}, L5 (2004);
730: C. O. Lousto and R. H. Price, Phys. Rev. D {\bf 69}, 087503
731: (2004).
732:
733: \bibitem{postNewt} A. G. Wiseman, Phys. Rev. D {\bf 46}, 1517 (1992);
734: L. E. Kidder, Phys. Rev. D {\bf 52}, 821 (1995);
735: L. Blanchet, M. S. S. Qusailah, and C. M. Will, Astrophys.
736: J. {\bf 635}, 508 (2005); T. Damour and A. Gopakumar, Phys. Rev. D {\bf 73}, 124006
737: (2006).
738:
739: \bibitem{closed} Z. Andrade and R. H. Price, Phys. Rev. D {\bf 56}, 6336
740: (1997); C. F. Sopuerta, N. Yunes, and P. Laguna, Phys. Rev. D
741: {\bf 74}, 124010 (2006); {\em Erratum-ibid.} {\bf 75} 069903 (2007);
742: Astrophys. J. {\bf 656}, L9 (2007).
743:
744: \bibitem{numrel}F. Pretorius,
745: Class. Quantum Grav. {\bf 22}, 425 (2005); Phys. Rev. Lett. {\bf 95}, 121101 (2005);
746: M. Campanelli, C. O. Lousto, P. Marronetti, and
747: Y. Zlochower, Phys. Rev. Lett. {\bf 96}, 111101 (2006);
748: J. G. Baker, J. Centrella, D.-I. Choi, M. Koppitz, and
749: J. van Meter, Phys. Rev. Lett. {\bf 96}, 111102 (2006).
750:
751: \bibitem{merger} J.A. Gonzalez, M. Hannam, U. Sperhake, B. Brugmann, and S. Husa,
752: Phys. Rev. Lett. {\bf 98}, 231101 (2007);
753: M. Campanelli, C.O. Lousto, Y. Zlochower, D. Merritt, Phys. Rev. Lett. {\bf 98}, 231102 (2007).
754:
755: \bibitem{consequences} P. Madau and
756: E. Quataert, Astrophys. J. {\bf 606}, L17 (2004);
757: Z. Haiman, Astrophys. J. {\bf 613} , 36 (2004);
758: A. Loeb, Phys. Rev. Lett. {\bf 99}, 041103 (2007);
759: A. Gualandris and D. Merritt,
760: Astrophys. J. {\bf 678}, 780 (2008);
761: S. Komossa and D. Merritt, Astrophys. J. {\bf 683}, L21 (2008);
762: L. Blecha and A. Loeb,
763: {\em Effects of gravitational-wave recoil on the dynamics
764: and growth of supermassive black holes}, arXiv:0805.1420.
765:
766: \bibitem{Kick} K. S. Thorne, Rev. Mod. Phys. {\bf 52}, 299 (1980);
767: L. E. Kidder, Phys. Rev. D {\bf 52}, 821 (1995);
768: M. Campanelli, C. O. Lousto, Y. Zlochower, and D. Merritt,
769: Astrophys. J. {\bf 659}, L5 (2007);
770: C. O. Lousto, Y. Zlochower, Phys. Rev. D {\bf 77}, 044028 (2008); C. O. Lousto, Y. Zlochower, {\em Modeling gravitational recoil from precessing highly-spinning unequal-mass black-hole binaries}, gr-qc/0805.0159;
771: S. H. Miller, R.A. Matzner, {\em Multipole Analysis of Kicks in Collision of Binary Black Holes}, arXiv:0807.3028.
772:
773: \bibitem{RT}I. Robinson and A. Trautman, Phys. Rev. Lett. {\bf 4}, 431
774: (1960); Proc. R. Soc. London A {\bf 265}, 463 (1962).
775:
776: \bibitem{book} H. Stephani, D. Kramer, M. MacCallum, C. Hoenselaers, E. Herlt,
777: {\em Exact Solutions of Einstein's Field Equations}, Second Edition,
778: Cambridge University Press, 2002.
779:
780: \bibitem{Chrusciel} P. Chrusciel, Comm. Math. Phys. {\bf 137}, 289 (1991)
781:
782: \bibitem{RTnum} R. Gomez, L. Lehner, P. Papadopoulos, and J Winicour,
783: Class. Quantum Grav. {\bf 14}, 977 (1997); O. Moreschi, A. Perez, and L. Lehner,
784: Phys. Rev. D {\bf 66}, 104017 (2002).
785:
786:
787: \bibitem{spectral} D. A. Prager and A. W.-C. Lun, J. Austral. Math. Soc. Ser. B {\bf 41}, 271 (1999).
788:
789: \bibitem{rio} H.P. de Oliveira and I. Dami\~ao Soares, Phys. Rev. D {\bf 70}, 084041 (2004).
790:
791: \bibitem{rio1} H.P. de Oliveira and I. Dami\~ao Soares, Phys. Rev. D {\bf 71}, 124034 (2005).
792:
793: \bibitem{galerkin} P. Holmes, J. L. Lumley, and G. Berkooz, {\em
794: Turbulence, Coherent Structures, Dynamical Systems
795: and Symmetry}, Cambridge University Press (1998).
796:
797: \bibitem{mass} R. K. Sachs, Phys. Rev. {\bf 128}, 2851 (1962);
798: H. Bondi, M. G. J. van der Berg, and A. W. K. Metzner, Proc. R. Soc. London A {\bf 269},
799: 21 (1962);
800: U. von der Goenna and D. Kramer, Class. Quant. Grav. {\bf 15}, 215 (1998).
801:
802:
803: \bibitem{hilbert} M.D. Choi, Amer. Math. Monthly {\bf 90}, 301 (1983).
804:
805: \bibitem{rio2} H.P. de Oliveira, I. Dami\~ao Soares, and E.V. Tonini, Phys. Rev. D {\bf 78}, 044016 (2008).
806:
807: \bibitem{Brill} D. R. Brill and R. W. Lindquist, Phys. Rev. {\bf 131}, 471 (1963).
808:
809: \bibitem{headone}O. M. Moreschi and S. Dain, Phys. Rev. D {\bf 53}, R1745 (1996).
810:
811:
812: \bibitem{headon} R.F. Aranha, H.P. de Oliveira, I. Dami\~ao Soares, and E.V. Tonini,
813: Int. J. Mod. Phys. {\bf D17}, 55 (2008).
814:
815:
816: \end{references}
817:
818:
819:
820:
821: \end{document}
822:
823:
824:
825:
826: