1:
2: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3: %
4: % $Id: id.tex,v 1.70 2003/05/19 14:55:17 bbh Exp $
5: %
6: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
7:
8: %\documentclass[prd, twocolumn, aps, showpacs]{revtex4}
9: \documentclass[prd, preprint, aps, showpacs]{revtex4}
10:
11: \usepackage{graphicx}
12: \usepackage{color}
13:
14: \def\Lie{\hbox{\it\char'44}}
15: \def\be{\begin{equation}}
16: \def\ee{\end{equation}}
17: \def\ba{\begin{eqnarray}}
18: \def\ea{\end{eqnarray}}
19: \def\beq{\begin{eqnarray}}
20: \def\eeq{\end{eqnarray} }
21:
22: \def\eref#1{Eq.~(\ref{#1})}
23: \def\Eref#1{Eq.~(\ref{#1})}
24: \def\fref#1{Fig.~\ref{#1}}
25: \def\Fref#1{Fig.~\ref{#1}}
26: \def\sref#1{Section~\ref{#1}}
27: \def\Sref#1{Section~\ref{#1}}
28: \def\tref#1{Table~\ref{#1}}
29: \def\Tref#1{Table~\ref{#1}}
30:
31: \def\Comment#1{{\textcolor{red}{\bf #1}}}
32: \def\BlueComment#1{{\textcolor{blue}{\bf #1}}}
33:
34: \def\rmd{{\rm d}}
35: \def\rmO{{\rm O}}
36: \def\ADM{{\rm ADM}}
37: \def\HOR{{\rm H}}
38: \def\ISCO{{\rm ISCO}}
39: \def\BE{{E_b}}
40: \def\lp{\left(}
41: \def\rp{\right)}
42: \def\lb{\left[}
43: \def\rb{\right]}
44: \def\etal{{et al.}}
45:
46:
47:
48: \begin{document}
49: \title{Physics and Initial Data for Multiple Black Hole Spacetimes}
50:
51: \author{Erin Bonning}
52: \author{Pedro Marronetti}
53: \altaffiliation{Department of Physics, University of Illinois at
54: Urbana-Champaign}
55: \author{David Neilsen}
56: \altaffiliation{Department of Physics \& Astronomy, Louisiana State University,
57: Baton Rouge LA 70803}
58: \author{Richard Matzner}
59: \affiliation{Center for Relativity, University of Texas at Austin,
60: Austin, TX 78712-1081, USA}
61:
62:
63:
64: \begin{abstract} An orbiting black hole binary will generate
65: strong gravitational radiation signatures, making these binaries important
66: candidates for detection in gravitational wave observatories.
67: The gravitational radiation is characterized by the orbital
68: parameters, including the frequency and separation at the inner-most
69: stable circular orbit (ISCO). One
70: approach to estimating these parameters relies on a sequence of initial
71: data slices that attempt to capture the physics of the inspiral. Using
72: calculations of the binding energy, several authors have estimated the
73: ISCO parameters using initial data constructed with various
74: algorithms. In this paper we examine this problem using conformally
75: Kerr-Schild initial data. We present convergence results for our
76: initial data solutions, and give data from numerical solutions of the
77: constraint equations representing a range of physical configurations.
78: In a first attempt to understand the physical content of the initial
79: data, we find that the Newtonian binding energy is contained in the
80: superposed Kerr-Schild background before the constraints are solved.
81: We examine some deficiencies with the initial data approach to orbiting
82: binaries, especially touching on the effects of prior motion and
83: spin-orbital coupling of the angular momenta. Making rough
84: estimates of these effects, we find that they are not insignificant
85: compared to the binding energy, leaving some doubt of the utility of
86: using initial data to predict ISCO parameters. In computations of
87: specific initial-data configurations we find spin-specific effects
88: that are consistent with analytical estimates.
89:
90:
91: \end{abstract}
92:
93: \pacs{ 02.30.Jr 02.40.Ky 02.60.Lj 02.70.Bf 04.20.Ex 04.25.Dm 95.30.Sf
94: 95.85.Sz}
95:
96: \maketitle
97:
98:
99: %----------------------------------------------------------------------
100: %
101: %
102: %
103: %----------------------------------------------------------------------
104: \section{Introduction}
105: \label{sec:intro}
106:
107:
108: The computation of gravitational wave production from the interaction
109: and merger of compact astrophysical objects is a
110: challenge which, when solved, will provide a predictive and analytical
111: resource for the upcoming gravitational wave detectors. A binary black
112: hole system is expected to be the strongest possible astrophysical
113: gravitational wave source. In particular, one expects a binary black
114: hole system to progress through a series of quasi-equilibrium states of
115: narrowing circular orbits as it emits gravitational radiation. In the
116: final moments of stellar mass black hole inspiral, the
117: radiation will be detectable in the current (LIGO-class) detectors. If
118: the total binary mass is of the order of $10M_{\odot}$, the moment of
119: final plunge to coalescence will emit a signal detectable by the
120: current generation of detectors from very distant (Gpc) sources.
121:
122: Detecting gravitational radiation is also a significant technical
123: challenge. Gravitational waves couple very weakly to matter, and the
124: expected signals are much smaller in amplitude than ambient environmental
125: and thermal noise. The successful detection of these waves, therefore,
126: requires some knowledge of what to look for.
127: In this regard, an orbiting binary black hole system is an ideal candidate
128: for detection since the orbital motion produces regular gravitational
129: radiation patterns. In such an inspiraling
130: black hole system, the strongest waves are emitted during the last
131: several orbits, as the holes reach the innermost quasi-stable orbit
132: (here abbreviated ISCO), and as they continue through the final
133: plunge. The dynamics of the holes during these final orbits, especially
134: the orbital angular velocity, $\omega_\ISCO$, and separation, $\ell_\ISCO$,
135: determine the dominant characteristics of the detectable waves.
136: Any knowledge of these parameters is advantageous for detecting
137: radiation from these binary systems.
138:
139: The proper way to predict gravitational waveforms for orbiting black
140: holes is to set initial data for two widely separated holes, and then
141: solve the evolution equations to follow the inspiral through merger and
142: beyond. This problem is well beyond the capabilities of current
143: evolution codes. Therefore, to obtain some information about orbiting black
144: holes we, and
145: others~\cite{Cook,Pfeiffer,Baumgarte,GGB1,GGB2,GGB3,Cook2,Pfeiffer2},
146: turn to the initial value
147: problem. For an introduction to the literature, see the review by
148: Cook~\cite{CookReview}.
149: Given a collection of initial data for black holes in circular
150: orbits with decreasing radius, one tries to identify a sequence of initial
151: data that corresponds to instantaneous images of a time-dependent
152: evolution. Circular orbits are chosen because orbits in the early
153: stages of an inspiral are predicted to become circularized
154: because of the stronger gravitational radiation near periapse~\cite{Peters}.
155: When a suitable sequence of initial data slices has been obtained, they can
156: then be used to determine various orbital parameters. For example, the
157: change in binding energy with respect to the orbital radius allows one
158: to identify $\ell_\ISCO$, and a similar analysis of the angular
159: momentum gives $\omega_{\rm ISCO}$. The difficulty in this approach
160: comes in ensuring that the initial data at one radius
161: correspond to the same physical system as the data for another
162: radius. This can be done for some systems by using
163: conserved quantities. For example, in the case of neutron stars, constant
164: baryon number is
165: an unambiguous indicator of the sameness of the stars. However, in black hole
166: physics it is not available; it is unclear how to determine that two black
167: hole initial data sets do, in fact, represent the same physical system.
168:
169: The initial data approach to studying binary black holes is thus not
170: without problems. These difficulties fall in two broad areas. First,
171: there is no unambiguous way to set initial data in general relativity.
172: The current algorithms all require some arbitrary
173: mathematical choices to find a solution. For instance, the approach we
174: take requires the definition of the topology of a background space and
175: of its metric and the momentum of the metric, followed by solution of four
176: coupled elliptic differential equations for variables that adjust the
177: background fields. But the choice of the background quantities is
178: arbitrary to a
179: large extent. The physical meaning of these mathematical choices is
180: not completely clear, but the effect is unmistakable. Data constructed
181: with various algorithms can differ substantially, even when attempting
182: to describe the same physical system~\cite{Pfeiffer2}. The data sets can be
183: demonstrated in many circumstances to contain the expected Newtonian
184: binding energy, as we show below (i.e., the binding energy of order
185: $\rmO(m^2/\ell)$ agrees with the Newtonian result at this order).
186: However, the data can differ significantly at $\rmO(m (m/\ell)^2)$.
187: These differences are attributed to differences in wave content of the
188: data which may reflect possible prior motion or may simply be
189: spurious. At present it is neither possible to build prior motion into
190: the initial data, nor to specify how radiation is added to the the
191: solution, nor to know how much there is. It is known that the
192: circular orbits and the ISCO so determined are in fact
193: method-dependent. Furthermore, the methods need not even agree that a
194: specific dataset represents a circular orbit; their subsequent
195: evolutions may not agree~\cite{Will}.
196:
197: A second problem---and the principal physical difficulty with the initial
198: data method for studying black
199: hole binaries---is the lack of unambiguous conserved quantities.
200: The best candidate for an invariant quantity is the event horizon area,
201: $A_{\rm H}$.
202: This area is unchanging for isentropic processes due to the proportionality
203: of $A_{\rm H}$ with the black hole entropy.
204: One can argue that since the quasi-circular orbit is quasi-adiabatic,
205: $A_{\rm H}$ is nearly invariant over some phase of the inspiral.
206: But the inspiral cannot be completely adiabatic because it cannot
207: be made arbitrarily slow; the black holes will absorb an
208: unknown amount of gravitational radiation while in orbit and will thereby
209: increase in size.
210: Moreover, the event horizon is a global construct of the spacetime,
211: and cannot be determined from a single slice of initial data.
212: Therefore, one must use the apparent horizon area, $A_{\rm AH}$,
213: as an {\it ersatz} invariant for initial data
214: studies~\cite{Shoemaker,Huq}.
215: When the hole is approximately stationary,
216: these horizon areas may be nearly equal~\cite{Dreyer}.
217: In dynamic configurations---as should be appropriate for orbiting
218: holes---these horizon areas may differ
219: substantially~\cite{Caveny,Caveny1}.
220:
221:
222: We will investigate physical content of initial data,
223: focusing on Kerr-Schild spacetimes. We examine binding energy to leading
224: order, and find that in our method of constructing
225: the superposed Kerr-Schild data, the background fields contain the
226: Newtonian binding energy: the subsequent solution of the elliptic equations
227: yields a only small correction. Using numerical solutions we present
228: orbital configurations with solved initial data. We give a qualitative
229: discussion of physical effects that may confound any attempt to study
230: inspiral via a sequence of initial data, and which may affect the
231: determination of the location of the ISCO. We give some computational
232: examples consistent with these qualitative predictions.
233:
234: %----------------------------------------------------------------------
235: %
236: %
237: %
238: %----------------------------------------------------------------------
239: \section{Review of initial data construction in general relativity}
240:
241: In the computational approach we take a Cauchy formulation
242: (3+1) of the ADM type, after Arnowitt, Deser, and
243: Misner~\cite{ADM}. In such a method the 3-metric $g_{ij}$ is the fundamental
244: variable. The 3-metric and its momentum are specified at one initial time on
245: a spacelike hypersurface. The ADM metric is
246: \be
247: \rmd s^2 = -(\alpha^2 - \beta_i \beta^i)\,\rmd t^2 + 2\beta_i \, \rmd t \,\rmd x^i
248: + g_{ij}\, \rmd x^i\, \rmd x^j
249: \label{eq:admMetric}
250: \ee
251: where $\alpha$ is the lapse function and $\beta^i$ is the shift 3-vector.
252: Latin indices run $1,2,3$ and are lowered and raised by $ g_{ij}$
253: and its 3-dimensional inverse $ g^{ij}$. $\alpha$ and $\beta^i$ are gauge
254: functions that relate the coordinates on each hypersurface to each other.
255: The extrinsic curvature, $K_{ij}$, plays the role of momentum conjugate
256: to the metric, and describes the embedding of a $t=\hbox{\rm constant}$
257: hypersurface into the 4-geometry.
258:
259: The Einstein field equations contain both hyperbolic evolution equations,
260: and elliptic constraint equations.
261: The constraint equations for vacuum in the ADM decomposition are:
262: \beq
263: R - K_{ij}K^{ij} + K^2 &=& 0,
264: \label{eq:constraintH}
265: \eeq
266: \beq
267: \nabla_j \lp K^{ij} - g^{ij}K\rp &=& 0.
268: \label{eq:constraintK}
269: \eeq
270:
271: Here $R$ is the 3-dimensional Ricci scalar, and $\nabla_j$ is the
272: 3-dimensional covariant derivative compatible with $ g_{ij}$. These
273: constraint equations guarantee a kind of transversality of the
274: momentum (\eref{eq:constraintK}). Initial data must satisfy these constraint
275: equations; one may not freely specify all components of $g_{ij}$ and
276: $K_{ij}$. The initial value problem in general relativity thus
277: requires one to consistently identify and separate constrained and
278: freely-specifiable parts of the initial data. Methods for making
279: this separation, and solving the constraints as an elliptic system,
280: include: the {\it conformal transverse-traceless
281: decomposition}~\cite{YP}; the {\it physical transverse-traceless
282: decomposition}~\cite{MY}: and the {\it conformal thin sandwich
283: decomposition} which assumes a helical killing
284: vector~\cite{Mathews, York,GGB1}.
285: These methods all involve
286: arbitrary choices and do not produce equivalent data. Our solution
287: method uses the conformal transverse-traceless decomposition~\cite{YP}.
288:
289: Solutions of the initial value problem have been addressed in
290: the past by several groups~\cite{Cook,YP,Baumgarte,Pfeiffer,GGB1}. It is the
291: case that until recently, most data have been constructed assuming that
292: the initial 3-space is conformally flat. The method most commonly used
293: is the approach of Bowen and York~\cite{Bowen+York}, which chooses
294: maximal spatial hypersurfaces and takes the spatial 3-metric to be
295: conformally flat. This method has been used to find candidate
296: quasi-circular orbits by Cook~\cite{Cook}, Baumgarte~\cite{Baumgarte}, and
297: most recently, Pfeiffer {\it et al.}~\cite{Pfeiffer}.
298:
299: The chief advantage of the maximal spatial hypersurface approach is
300: numerical simplicity, as the choice $K = 0$ decouples the Hamiltonian
301: constraint from the momentum constraint equations. If, besides $K =
302: 0$, the conformal background is flat Euclidean 3-space, there are known
303: $K_{ij}$ that analytically solve the momentum
304: constraint~\cite{Bowen+York}. The constraints then reduce to one
305: elliptic equation for the conformal factor $\phi$. However, it has
306: been pointed out by Garat and Price~\cite{Garat} that there are no
307: conformally flat $K=0$ slices of the Kerr spacetime. Since we expect
308: astrophysical sources to be rotating, the choice of a conformally flat
309: $K=0$ background will yield data that necessarily contains some
310: quantity of ``junk" gravitational radiation. Jansen {\it et al.}~\cite{Jansen}
311: have recently shown by comparison with known solutions that conformally flat
312: data do indeed contain a significant amount of unphysical gravitational
313: field. Another conformally flat $K=0$ method recently
314: used by Gourgoulhon, Grandclement, and
315: Bonazzola~\cite{GGB1,GGB2} is a thin sandwich approximation based on the
316: approach of Wilson and Mathews~\cite{Mathews} which assumes the presence of an
317: instantaneous rotation Killing vector to define the initial extrinsic
318: curvature. They impose a specific gauge defined by demanding that $K$ and the
319: conformal factor remain constant in the rotating frame.
320: Since $\phi$ and $K$ are a conjugate pair
321: in the ADM approach, this method solves the four initial value equations and
322: one second-order evolution equation. The assumption of a Killing
323: vector suppresses radiation or, perhaps more accurately, imposes a
324: condition of equal ingoing and outgoing radiation.
325:
326: In this paper we use Kerr-Schild data~\cite{Matzner:1999pt} to outline some
327: of the difficulties in
328: finding the ISCO using the initial data technique. We discuss the extent
329: to which initial data set by means of superposed Kerr-Schild black holes limits
330: the extraneous radiation in the data sets, and we estimate the accuracy
331: of the extant published ISCO determinations.
332: Recent works by Pfeiffer, Cook, and Teukolsky also investigate binary black
333: hole systems using Kerr-Schild initial data~\cite{Pfeiffer2}.
334:
335:
336: %----------------------------------------------------------------------
337: %
338: %
339: %
340: %----------------------------------------------------------------------
341: \section{Initial Data \lowercase{via} Superposed Kerr-Schild Black Holes}
342: \label{sec:ks_id}
343:
344:
345: The superposed Kerr-Schild method for setting black hole initial data,
346: developed by Matzner, Huq, and Shoemaker~\cite{Matzner:1999pt},
347: produces data for black holes of arbitrary masses, boosts, and spins
348: without relying on any underlying symmetries of any particular
349: configuration. The method proceeds in two parts. First, a background
350: metric and background extrinsic curvature are constructed by
351: superposing individual Kerr-Schild black hole solutions. Then the
352: physical data are generated by solving the four coupled constraint
353: equations for corrections to the background. Intuitively, the
354: background solution should be very close to the genuine solution when
355: the black holes are widely separated, and only small adjustments to the
356: gravitational fields are required to solve the constraints. We show
357: that this is true for large and also for small separations. This
358: section briefly reviews the superposed Kerr-Schild method for initial
359: data, then gives some analytic results to justify this contention.
360:
361: %----------------------------------------------------------------------
362: %
363: %
364: %
365: %----------------------------------------------------------------------
366: \subsection{Kerr-Schild data for isolated black holes}
367: \label{sec:ks}
368:
369: The Kerr-Schild~\cite{KerrSchild} form of a black hole solution describes the
370: spacetime of a single black hole
371: with mass, $m$, and specific angular momentum, $a = j/m$, in a coordinate
372: system that is well behaved at the horizon.
373: (We use uppercase $M$ for calculated masses, e.g., the ADM mass,
374: and lowercase $m$ for mass parameters, or when the distinction
375: is not important.)
376: The Kerr-Schild metric is
377: \be
378: \rmd s^{2} = \eta_{\mu \nu}\,\rmd x^{\mu}\, \rmd x^{\nu}
379: + 2H(x^{\alpha}) l_{\mu} l_{\nu}\,\rmd x^{\mu}\,\rmd x^{\nu},
380: \label{eq:1}
381: \ee
382: where $\eta_{\mu \nu}$ is the metric of flat space, $H$ is a scalar
383: function of $x^\mu$, and $l_{\mu}$ is an (ingoing) null vector, null
384: with respect to both the background and the full metric,
385: \be
386: \eta^{\mu \nu} l_{\mu} l_{\nu} = g^{\mu \nu} l_{\mu} l_{\nu} = 0.
387: \label{eq:2}
388: \ee
389: This last condition gives $l^0 l_0 = - l^i l_i$.
390:
391: The general non-moving
392: black hole metric in Kerr-Schild form (written in Kerr's original
393: rectangular coordinates) has
394: \begin{equation}
395: H = \frac{mr}{r^{2} + a^{2}\cos^{2} \theta},
396: \label{eq:ks_h}
397: \end{equation}
398: and
399: \begin{equation}
400: l_{\mu} = \left(1, \frac{rx + ay}{r^{2} + a^{2}}, \frac{ry -
401: ax}{r^{2} + a^{2}}, \frac{z}{r}\right),
402: \label{eq:4}
403: \end{equation}
404: where $r,~ \theta$ (and $\phi$)
405: are auxiliary spheroidal coordinates, $z = r(x,y,z) \cos \theta$,
406: and $\phi$ is
407: the axial angle. $r(x, y, z)$ is obtained from the relation,
408: \begin{equation}
409: \frac{x^{2} + y^{2}}{r^{2} + a^{2}} + \frac{z^{2}}{r^{2}} = 1,
410: \label{eq:5}
411: \end{equation}
412: giving
413: \begin{equation}
414: r^{2} = \frac{1}{2}(\rho^{2} - a^{2}) +
415: \sqrt{\frac{1}{4}(\rho^{2} - a^{2})^{2} + a^{2}z^{2}},
416: \label{eq:6}
417: \end{equation}
418: with
419: \begin{equation}
420: \rho = \sqrt{x^{2} + y^{2} + z^{2}}.
421: \label{eq:rho_def}
422: \end{equation}
423:
424: Comparing the Kerr-Schild metric with the ADM
425: decomposition~\eref{eq:admMetric}, we find that the $t=\hbox{\rm constant}$
426: 3-space metric is:
427:
428: \be
429: g_{ij} = \delta_{ij} + 2 H l_i l_j,
430: \label{eq:3metric_ks}
431: \ee
432:
433: Further, the ADM gauge variables are
434: \be
435: \beta_i = 2 H l_0 l_i,
436: \label{eq:beta_ks}
437: \ee
438: and
439: \be
440: \alpha = \frac{1}{\sqrt{1 + 2 H l_0^2}}.
441: \ee
442:
443: The extrinsic curvature can be computed from the metric using the ADM
444: evolution equation~\cite{MTW}
445: \be
446: K_{ij} = \frac{1}{2\alpha}[\nabla_j\beta_{i} + \nabla_i\beta_{j}
447: - \dot g_{ij}],
448: \label{eq:k_ks}
449: \ee
450: where a dot ( $\dot{}$ ) denotes a the partial derivative with respect
451: to time.
452: Each term on the right hand side of this equation is known analytically.
453:
454:
455:
456: %----------------------------------------------------------------------
457: %
458: %
459: %
460: %----------------------------------------------------------------------
461: \subsection{Boosted Kerr-Schild black holes}
462:
463: The Kerr-Schild metric is form-invariant under a
464: boost, making it an ideal metric to describe moving
465: black holes. A constant Lorentz transformation
466: (the boost velocity, ${\bf v}$, is specified with respect to the background
467: Minkowski spacetime) $\Lambda^{\alpha}{}_{\beta}$ leaves the
468: 4-metric in Kerr-Schild form, with $H$ and $l_{\mu}$
469: transformed in the usual manner:\\
470: \ba
471: x'^{\beta} &=& \Lambda^\beta{}_\alpha x^{\alpha},\\
472: H'(x'^{\alpha}) &=& H\lp (\Lambda^{-1})^\alpha{}_\beta
473: \,\,x'^{\beta}\rp,\\
474: l'_{\delta}(x'^{\alpha}) &=& \Lambda^{\gamma}{}_{\delta}\,\,
475: l_{\gamma}\lp(\Lambda^{-1})^\alpha{}_\beta\,\, x'^{\beta}\rp .
476: \label{eq:ks_boost}
477: \ea
478: Note that $l'_{0}$ is no longer unity. As the initial solution
479: is stationary, the only time dependence comes in the
480: motion of the center, and the full metric is stationary with a Killing
481: vector reflecting the boost velocity.
482: The solution, therefore, contains no junk radiation, as no radiation
483: escapes to infinity during a subsequent evolution.
484: Thus, Kerr-Schild data exactly represent a spinning and/or moving single
485: black hole. This is not possible in some other approaches, e.g.,
486: the conformally flat approach~\cite{Jansen}.
487:
488: %----------------------------------------------------------------------
489: %
490: %
491: %
492: %----------------------------------------------------------------------
493: \subsection{Background data for multiple black holes}
494:
495: The structure of the Kerr-Schild metric suggests a natural extension
496: for multiple black hole spacetimes using the straightforward superposition
497: of flat space and black hole functions
498: \be
499: {g}_{ij} \approx \eta_{ij}
500: + 2~{}_1H_{~1}l_{i~1}l_{j}
501: + 2~{}_2H_{~2}l_{i~2}l_{j}
502: + \cdots\, ,
503: \label{eq:metric_super_simple}
504: \ee
505: where the preceding subscript numbers the black holes. Note that a
506: simple superposition is typically {\em not} a genuine solution of the Einstein
507: equations, as it does not
508: satisfy the constraints, but it should be ``close'' to the real solution
509: when the holes are widely separated.
510:
511: To generate the background data,
512: we first choose mass and angular momentum parameters for each hole,
513: and compute the respective $H$ and $l^\alpha$ in the appropriate
514: rest frame. These quantities are then boosted in the desired direction
515: and offset to the chosen position in the computational frame.
516: The computational grid is the center of momentum frame for the two holes,
517: making the velocity of the second hole a function of the two
518: masses and the velocity of the first hole.
519: Finally, we compute the
520: individual metrics and extrinsic curvatures in the coordinate system
521: of the computational domain:
522: \beq
523: {}_A g_{ij} &=& \eta_{ij}
524: + 2~{}_A H ~{}_A l_{i} ~{}_A l_{j},\\
525: {}_A K_i{}^m &=& \frac{1}{2\alpha} ~{}_A g^{mj}
526: \lp \nabla_j ~{}_A\beta_{i} + \nabla_i~{}_A \beta_{j}
527: - ~{}_A \dot g_{ij}\rp.
528: \eeq
529: %
530: Again, the index $A$ labels the black holes.
531: Data for $N$ holes are then constructed in superposition
532: %\beq
533: %\tilde{g}_{ij} &=& \eta_{ij} + 2~{}_{1}B{}_{~1}H_{~1}l_{i~1}l_{j}
534: % + 2~{}_{2}B{}_{~2}H_{~2}l_{i~2}l_{j}\,,\nonumber\\
535: %\tilde{K} &=& {}_{1}B{}_{~1}K_i^{~i}+{}_{2}B{}_{~2}K_i^{~i}\,,\\
536: %\tilde{A}_{ij} &=& \tilde{g}_{n(i}~~({}_{1}B{}_{~1}K_{j)}^{~n}
537: % +{}_{2}B{}_{~2}K_{j)}^{~n}
538: % - \frac{1}{3} \delta_{j)}^{~n} \tilde{K})\, .\nonumber
539: %\label{eq:ks_super}
540: %\eeq
541: %
542:
543: \beq
544: \tilde{g}_{ij} &=& \eta_{ij} + \sum_A^N 2~{}_AB~{}_A H {}_A l_i ~{}_A l_j ,\\
545: \tilde{K} &=& \sum_A^N {}_AB~{}_AK_i{}^i ,\\
546: %\tilde{A}_{ij} &=& \tilde{g}_{n(i}~~\sum_A^N \lp {}_AB~{}_AK_{j)}{}^n
547: % -\frac{1}{3} \delta_{j)}{}^n {}_AB~{}_AK_i{}^i\rp .
548: \tilde{A}_{ij} &=& \tilde{g}_{n(i}~~\sum_A^N {}_AB~\lp {}_AK_{j)}{}^n
549: -\frac{1}{3} \delta_{j)}{}^n ~{}_AK_i{}^i\rp .
550: \label{eq:ks_super}
551: \eeq
552: %
553: A tilde ( $\tilde{}$ ) indicates a background field tensor.
554: The simple superposition of the metric from~\eref{eq:metric_super_simple}
555: (part of the original specification~\cite{Matzner:1999pt}) has been modified
556: here with the introduction of
557: {\it attenuation functions}, ${}_A B$~\cite{Marronetti,Marronetti:2000rw}.
558: The extrinsic curvature is separated into its trace, $K$, and trace-free
559: parts, $A_{ij}$, and the indices of $\tilde{A}_{ij}$ are
560: explicitly symmeterized.
561:
562: The attenuation functions represent the physical idea that in the immediate
563: vicinity of one hole, the effect of a second hole becomes negligible.
564: Near a black hole the conformal background
565: superposition ( $ \tilde{} $ ) metrics approach the analytic values for
566: the single black hole.
567: The attenuation
568: function ${}_{2}B$ (${}_1B$) eliminates the
569: influence of the second (first) black hole in the vicinity of the first
570: (second).
571: ${}_{1}B$ equals unity everywhere except in the vicinity of the second
572: black hole, and its first and second derivatives are zero at the
573: singularity of the second hole.
574:
575: The attenuation function used is
576:
577: \be
578: {}_{1}B =1-\exp(-\ell_{1}^{4}/2\sigma^2),
579: \ee
580: where $\ell_1$ is the
581: coordinate distance from the center of hole $2$,
582:
583: \ba
584: \ell_1^{2} &=& \frac{1}{2}(\rho^{2} - a^{2}) +
585: \sqrt{\frac{1}{4}(\rho^{2} - a^{2})^2 + a^{2}z^{2}}~~,\\
586: \rho &=& \sqrt{{}_{2}\gamma^2(x-{}_{2}x)^{2} + (y-{}_{2}y)^{2}
587: + (z-{}_{2}z)^{2}}~.
588: \ea
589:
590: %
591: \noindent and $\sigma$ is a parameter. In all examples given in this paper,
592: the masses are equal and $\sigma = m^2$.
593: \Fref{fig:attenuation} shows a typical attenuation function used in
594: calculating our initial data sets.
595:
596: \begin{figure}
597: \begin{center}
598: \includegraphics[width=5.in]{fig1.ps}
599: \end{center}
600: \caption{
601: The attenuation function,
602: ${}_1 B =1-\exp(-\ell_{1}^{4}/2\sigma^2)$, used to calculate our initial
603: data solutions.
604: To indicate the effect of the attenuation function in a binary black
605: hole system, we also plot the
606: the background metric function $\tilde g_{yy}$
607: in the vicinity of one hole with and without attenuation.
608: The Schwarzschild black holes are placed along the $y$-axis at $\pm 4 m$.
609: Here $\ell_1$ is the coordinate distance from the center
610: of the second black hole, and the attenuation function width is
611: $\sigma = m^2$.
612: }
613: \label{fig:attenuation}
614: \end{figure}
615:
616: A small volume containing the singularity
617: is masked from the computational domain.
618: This volume is
619: specified by choosing a threshold value for the Ricci scalar, typically
620: for $|R| \geq 2/m^2$. For a single Schwarzschild black hole, this gives a
621: spherical mask with a radius $r \simeq 0.73~m$. In all cases the masked
622: region lies well within apparent horizons in the solved data. In
623: practice we find that a small attenuation region (also inside the
624: apparent horizon) is necessary to achieve a smooth solution of the
625: elliptic initial data equations near the mask; see Section {\bf II.D} below.
626: Figures \ref{fig:hc_with_att} and
627: \ref{fig:mc_with_att} show the Hamiltonian and momentum constraints for
628: the background space with and without attenuation. We have not varied
629: the masking condition to determine what effect the size of the mask has
630: on the global solution. As mentioned below, Pfeiffer {\it et al.}
631: have investigated this point~\cite{Pfeiffer2}.
632:
633: \begin{figure}
634: \begin{center}
635: \includegraphics[width=3.5in,angle=270]{fig2.ps}
636: \end{center}
637: \caption{The Hamiltonian constraint (units $m^{-2}$) calculated for the
638: background space for two identical Schwarzschild black holes.
639: The black holes are located on the $y$-axis at $y = \pm 4$~m,
640: and have zero initial velocity.
641: The solid curve is the background behavior of the constraint
642: without using attenuation functions, and the dashed curve is the
643: constraint
644: with attenuation and $\sigma = m^2$. The masked region is within the
645: radius $r ~~\widetilde{<}~~0.73 m$. It can be seen that attenuation
646: does not necessarily reduce the constraint, but does smooth it.}
647: \label{fig:hc_with_att}
648: \end{figure}
649:
650: \begin{figure}
651: \begin{center}
652: \includegraphics[width=3.5in,angle=270]{fig3.ps}
653: \end{center}
654: \caption{The $y$-component of the momentum constraint (units $m^{-2}$)
655: calculated for the background space of two identical Schwarzschild
656: black holes. The black holes are located on the $y$-axis at $y = \pm 4~m$,
657: and have zero initial velocity. The solid curve is the background
658: behavior
659: of the constraint
660: without using attenuation functions, and the dashed curve is the
661: constraint
662: with attenuation and $\sigma = m^2$.
663: }
664: \label{fig:mc_with_att}
665: \end{figure}
666:
667:
668:
669: %----------------------------------------------------------------------
670: %
671: %
672: %
673: %----------------------------------------------------------------------
674: \subsection{Generating the physical spacetime}
675:
676: The superposition of Kerr-Schild data described in the previous section
677: does not satisfy the constraints,
678: Eqs.~(\ref{eq:constraintH})--(\ref{eq:constraintK}), and hence are not
679: physical.
680: A physical spacetime can be constructed by modifying the background
681: fields with new functions such that the constraints are satisfied.
682: We adopt the conformal transverse-traceless method
683: of York and collaborators~\cite{YP} which consists of a
684: conformal decomposition
685: and a vector potential that adjusts the longitudinal components of the
686: extrinsic curvature.
687: The constraint equations are then solved for these new quantities such that
688: the complete solution fully satisfies the constraints.
689:
690: The physical metric, $g_{ij}$, and the trace-free part of the extrinsic
691: curvature, $A_{ij}$, are related to the background fields through a conformal
692: factor
693: \ba
694: g_{ij} &=& \phi^{4} \tilde{g}_{ij}, \label{confg1} \\
695: \label{confg}
696: A^{ij} &=& \phi^{-10} (\tilde{A}^{ij} + \tilde{(lw)}^{ij}),
697: \label{eq:conf_field}
698: \ea
699: where $\phi$ is the conformal factor, and $\tilde{(lw)}^{ij}$
700: will be used to cancel any possible longitudinal contribution to the
701: superposed background extrinsic curvature.
702: $w^i$ is a vector potential, and
703: \ba
704: \tilde{(lw)}^{ij} \equiv \tilde{\nabla}^{i} w^{j} + \tilde{\nabla}^{j} w^{i}
705: - \frac{2}{3} \tilde{g}^{ij} \tilde{\nabla_{k}} w^{k}.
706: \label{lw}
707: \ea
708: The trace $K$ is taken to be a given function
709: \be
710: K = \tilde K.
711: \label{tk}
712: \ee
713: Writing the Hamiltonian and momentum constraint equations in terms of
714: the quantities in
715: Eqs.~(\ref{confg1})--(\ref{tk}), we obtain four coupled
716: elliptic equations for the fields $\phi$ and $w^i$~\cite{YP}:
717: \ba
718: \tilde{\nabla}^2 \phi &=& (1/8) \big( \tilde{R}\phi
719: + \frac{2}{3} \tilde{K}^{2}\phi^{5} - \nonumber \\
720: & & \phi^{-7} (\tilde{A}{^{ij}} + (\tilde{lw})^{ij})
721: (\tilde{A}_{ij} + (\tilde{lw})_{ij}) \big), \\
722: \tilde{\nabla}_{j}(\tilde{lw})^{ij} &=& \frac{2}{3} \tilde{g}^{ij} \phi^{6}
723: \tilde{\nabla}_{j} K - \tilde{\nabla}_{j} \tilde{A}{^{ij}}.
724: \label{ell_eqs}
725: \ea
726:
727:
728: \subsection{Boundary Conditions}
729: \label{sec:boundary}
730:
731: A solution of the elliptic constraint equations requires that boundary
732: data be specified on both the outer boundary {\em and } the surfaces
733: of the masked regions. This contrasts with the hyperbolic evolution
734: equations for which excision can in principle be carried out without
735: setting inner boundary data since no information can propagate
736: out of the holes. Boundaries in an elliptic system, on the other hand,
737: have an immediate influence on the entire solution domain.
738: Using the attenuation functions, we can choose simple
739: conditions, $\phi = 1$ and $w^{i} = 0$, on the masked regions surrounding
740: the singularities.
741: In practice this inner boundary condition is not completely satisfactory
742: because it generates small discontinuities in the solution at this boundary.
743: These discontinuities are small relative to the scales in the problem,
744: and are contained within the horizon.
745: We have made no attempt to determine their global effect on the solution.
746: Pfeiffer \etal~\cite{Pfeiffer2} report a similar observation,
747: and note that the location of the boundary does affect some
748: aspects of the solution, though it has little effect on the fractional binding
749: energy or the location of the ISCO.
750:
751: The outer boundary conditions are more interesting. Several physical
752: quantities of interest, e.g., the ADM mass and momenta,
753: are global properties of the spacetime, and are calculated
754: on surfaces near the outer boundary of the computational
755: grid. Hence the outer boundary conditions must be chosen carefully
756: to obtain the proper physics.
757: We base our outer boundary conditions on an asymptotic expansion of the
758: Kerr-Schild metric, which relies on the ADM
759: mass and momentum formul\ae\ to identify the physically relevant terms
760: at the boundaries. We first review these expansions and
761: formul\ae.
762:
763: An asymptotic expansion of the Kerr-Schild metric ($\rho \gg m$)
764: gives
765: \beq
766: r &=& \rho \left(1 + \rmO (\rho^{-2})\right),\\
767: %
768: H &=& m/\rho \left( 1 + \rmO (\rho^{-2}) \right),\\
769: \label{eq:h_inf}
770: l_{i} &=& n_{i} + \frac{a^{c}\epsilon_{ijc} n^{j}}{\rho} + \rmO (
771: \rho^{-2} ),
772: \label{eq:l_inf}
773: \eeq
774: %
775: where
776: %
777: $n_{i} = n^{i} = x^{i}/\rho$.
778: (This is the only place where we do {\it not} use the 3-metric to raise
779: and lower indices, and $n_i n^i = 1$).
780: $a^{c}$ is the Kerr spin parameter with a general direction:
781: $a^{c} = a \hat a^{c}$.
782: %
783: The shift (Eq.~\ref{eq:beta_ks}) is asymptotically
784: \begin{equation}
785: \beta_{i} = \frac{2m}{\rho} \left( n_{i} + a^{c}
786: \frac{\epsilon_{ijc}n^{j}}{\rho} \right) + \rmO(\rho^{- 3}).
787: \label{eq:beta_inf}
788: \end{equation}
789: %
790: The asymptotic expansion of the extrinsic curvature in the stationary
791: Kerr-Schild form (cf.\ \eref{eq:k_ks}) is
792: %
793: \begin{eqnarray}
794: \alpha K_{ab} &=&
795: \frac{2m}{\rho^{2}} \left( -2n_{a}n_{b} + \delta_{ab}\right) \nonumber\\
796: & & \hspace{0.1cm} - \frac{3m}{\rho^{3}} a^{c}
797: \left( \epsilon_{ajc}n_{b} + \epsilon_{bjc}n_{a}\right)n_{j}
798: \nonumber\\
799: & & \hspace{0.1cm}
800: + \frac{6m^2}{\rho^{3}} \left(n_{a}n_{b} -\frac{2}{3} \delta_{ab}\right)
801: + \rmO(\rho^{-4}) .
802: \label{eq:k_inf}
803: \end{eqnarray}
804: The terms proportional to $a^c/\rho^3$ in this expression arise from the
805: transverse components of $\beta^a$ ($\beta_a n^a = 0$); the terms
806: of $ \rmO(\rho^{-3})$ independent of
807: $a^c$ arise from the affine connection.
808: Note that $\alpha= 1+ \rmO(\rho^{-1})$, and will
809: not affect the ADM estimates below.
810:
811:
812: The ADM formul\ae\ are evaluated in an asymptotically flat region
813: surrounding the system of interest, and in Cartesian coordinates they are
814: \beq
815: \label{eq:adm_mass}
816: M_{\ADM} &=& \frac{1}{16\pi} \oint \left( \frac{\partial g_{ji}}{\partial
817: x^{j}} - \frac{\partial g_{jj}}{\partial x^{i}} \right)
818: \rmd S^i,\\
819: \label{eq:adm_mom}
820: P^{\ADM}_{k} &=& \frac{1}{8\pi} \oint \left( K_{ki} - K^{b}{}_{b} \delta_{ki}
821: \right)\rmd S^i,\\
822: \label{eq:adm_ang_mom}
823: J^{\ADM}_{ab} &=& \frac{1}{8\pi} \oint \left( x_{a}K_{bi} - x_{b}K_{ai}
824: \right) \rmd S^i,
825: \eeq
826: for the mass, linear momentum, and angular momentum of the system
827: respectively~\cite{Wald, note4}. (All repeated indices are summed.)
828: The mass and linear momentum together constitute a 4-vector under Lorentz
829: transformations in the asymptotic Minkowski space, and
830: the angular momentum depends only on the trace-free
831: components of the extrinsic curvature.
832:
833: To compute the ADM mass and moment\ae for a single, stationary Kerr-Schild
834: black hole, we evaluate the integrals on the surface of a distant sphere.
835: The surface element then becomes
836: $dS^i = n^i \rho^2 \rmd\Omega$, where $n^i$ is the outward normal and
837: $\rmd\Omega$ is the differential solid angle.
838: we need to evaluate the metric only to order $\rmO(\rho^{-1})$;
839: the differentiation in Eq.~(\ref{eq:adm_mass}) guarantees that
840: terms falling off faster than $\rho^{-1}$ do not contribute to the
841: integration. The integrand is then
842: %
843: $ \frac{4m}{\rho^{2}} n_{i} \rho^{2}n^i \,\rmd\Omega$
844: %
845: and the integration yields the expected ADM mass $M_{ADM}=m$.
846: The ADM linear momentum requires only the leading order of of $K_{ab}$,
847: $\rmO(\rho^{-2})$;
848: terms falling off faster than this do not contribute.
849: The integrand of \eref{eq:adm_mom} then becomes
850: %
851: $-\frac{4m}{\rho^{2}} n_{a}n_{b} n^{b} \rho^{2} \,\rmd\Omega,$
852: %
853: yielding zero for the 3-momentum, as expected for a non-moving black
854: hole.
855:
856: At first blush, the integral for the ADM angular
857: momentum~\eref{eq:adm_ang_mom} appears
858: warrant some concern:
859: To leading order $K_{ab}$ is $\rmO({\rho^{-2}})$,
860: and the explicit appearance of $x_{a}$ in the integrand suggests that it
861: grows at infinity as $\rmO(\rho)$,
862: leading to a divergent result.
863: %
864: However, inserting the leading order term of $K_{ab}$ for a single,
865: stationary Kerr-Schild black hole into the integrand of
866: \eref{eq:adm_ang_mom}, we find that the integrand is identically zero.
867: The $\rmO({\rho^{-2}})$ terms of $K_{ab}$ contain the quantities
868: $n_a n_b$ and $\delta_{ab}$, which separately cancel because of the
869: antisymmetric form of \eref{eq:adm_ang_mom},
870: %
871: %
872: and a divergent angular momentum is avoided.
873: Including the $\rmO(\rho^{-3})$ terms of $K_{ab}$, we find
874: $J^\ADM_{ab} = \epsilon_{abc}a^{c}m$; the symmetry of the other
875: $\rmO(m^2\rho^{-3})$ terms again means they do not contribute.
876: This result for $J^\ADM_{ab}$ thus depends on
877: terms in the integrand proportional to $a$ that arise from corresponding
878: terms in $\beta^i$ proportional to $q^i$ where
879: $q^a$ is a unit vector transverse to the radial direction, $q^a n_a=0$.
880: Only these terms contribute to
881: the angular momentum integral; in particular those terms in $\beta^i$
882: proportional to
883: $n^i/\rho$ do not contribute.
884:
885: The ADM mass and momenta are Lorentz invariant. For a single,
886: boosted black hole, we naturally obtain $M_\ADM = \gamma m$ and
887: $P_\ADM = \gamma m v$.
888: The background spacetime for multiple black holes is constructed
889: with a superposition principle, and the ADM quantities are linear in
890: deviations about flat space at infinity.
891: Thus the ADM formul\ae, evaluated at infinity in the superposition,
892: {\it do} yield the expected superposition. For example, given two
893: widely separated
894: black holes boosted in the $x$-$y$ plane with spins aligned along
895: the $z$-axis, we have
896: \beq
897: \tilde M_\ADM &=&{}_1 \gamma {}_1 m + {}_2\gamma {}_2 m,\\
898: \tilde P^\ADM_i &=& 0, \\
899: \tilde J^\ADM_{12} &=& {}_1\gamma \left({}_1m {}_1v {}_1b
900: + {}_1m {}_1a\right) + {}_2\gamma \left({}_2m {}_2v {}_2b
901: + {}_2m {}_2a\right),
902: \label{eq:boosted}
903: \eeq
904: where ${}_1b$ and ${}_2b$ are impact parameters~\cite{note1},
905: and the tilde ( $ \tilde{} $ ) superscript
906: indicates that these quantities are calculated with the background tensors
907: $\tilde g_{ab}$ and $\tilde K_{ab}$. This superposition principle for
908: the ADM quantities in the background data is one advantage of conformal
909: Kerr-Schild initial data. (Note, in choosing the center of
910: momentum frame for the computation, $P^{\ADM}_{i} =0$ is a condition for
911: setting the background data.)
912:
913: Consider now the ADM integrals for the solved data. The Hamiltonian
914: constraint becomes an equation for the conformal factor, $\phi$.
915: As this equation is a nonlinear generalization of Poisson's equation,
916: asymptotic flatness in the full, solved metric requires that
917: \be
918: \phi \longrightarrow 1 + \frac{C}{2\rho} + \rmO(\rho^{-2}),
919: \ee
920: where $C$ is a (finite) constant.
921: This leads to our outer boundary condition for $\phi$, namely
922: \begin{equation}
923: \partial_{\rho} \left( \rho (\phi - 1) \right)|_{\rho \rightarrow
924: \infty} = 0.
925: \label{eq:phi_boundary}
926: \end{equation}
927: Furthermore, the linearity of the ADM mass integral gives
928: \begin{equation}
929: M_\ADM\mbox{(solved)} = {}_1\gamma{}_{1}m +
930: {}_2\gamma {}_{2}m + C.
931: \label{eq:m_adm_solved}
932: \end{equation}
933: (Here the absence of a tilde ( $ \tilde{} $ )
934: indicates that this mass is calculated
935: using the solved $g_{ab}$.) At this point we cannot predict even the
936: sign of $C$, though $|C|$ is expected to be small for widely separated
937: holes. If $|C|\to \infty$, then the boundary condition
938: \eref{eq:phi_boundary} would fail. The existence of solutions using
939: this condition, however, provides evidence that this possibility does not
940: occur.
941:
942: The boundary condition for $w^{i}$ is more subtle. {\it A priori}, we
943: expect $w^{i} \rightarrow 0$ at infinity, but a physically
944: correct solution on a finite domain requires that we understand how
945: $w^{i}$ approaches this limit at infinity.
946: We construct our boundary conditions on $w^k$ by demanding that the
947: ADM angular momentum of the full (solved) system be only finitely
948: different from that of the background (superposed) data.
949: That is, given that $\{\tilde g_{ab}, \tilde K_{ab}\}$ and
950: $\{g_{ab},K_{ab}\}$ have finite differences at infinity, we demand that
951: $J_{ab} -\tilde J_{ab}$ also be finite.
952: Using (\eref{eq:conf_field}) and (\eref{eq:adm_ang_mom}),
953: we find for the difference in angular momentum
954: \be
955: J_{ab}- \tilde J_{ab} = \frac{1}{8\pi} \oint\left( x_{a}
956: \nabla_{(b} w_{i)} - x_{b}
957: \nabla_{(a} w_{i)} \right)\rmd S^i\,.
958: \label{eq:j_adm_diff}
959: \ee
960: ($\phi\to 1$ at infinity, and there is
961: no difference at this order between conformal and physical versions of
962: $w^i$ and $g_{ab}$
963: at infinity.)
964:
965: We have already evaluated an integral of this form, in the discussion
966: of the Kerr angular momentum (see \eref{eq:adm_ang_mom} and \eref{eq:k_inf}),
967: where we expressed $K_{ab}$ in terms of the Kerr-Schild shift vector.
968: In that analysis, we noted that falloff of the form
969: \be
970: w_{i} \longrightarrow \frac{C_1}{\rho}n_{i} + \frac{C_2}{\rho^{2}}
971: q_{i} + \rmO(\rho^{-3}),
972: \label{24}
973: \ee
974: with $C_1$ and $C_2$ constant, and $q_i n^i = 0$,
975: will give a finite contribution to the angular momentum.
976: We therefore take as boundary conditions:
977: \begin{eqnarray}
978: & & \partial_\rho (\rho w^{i} n_{i}) = 0 \\[.12in]
979: \label{25}
980: & & \partial_\rho \left( \rho^{2} w^{i} (\delta_{ij} - n_{i}n_{j})
981: \right) = 0\,.
982: \label{26}
983: \end{eqnarray}
984:
985: Figures (\ref{fig:circular_orbit_nospin_1phi})--(\ref{fig:circular_orbit_nospin_1wy})
986: display $\phi$ and $w^i$ for a simple configuration. In this case the
987: elliptic equations were solved on a domain of $\pm 10~m$ along each axis
988: with resolution $\Delta x = m/8$.
989: As can be seen in these figures, the functions $\phi$ and $w^i$ actually
990: result in little adjustment to the background configurations.
991: Also note that the radial component of $w^i$, $w^i n_i$,
992: is the dominant function. In the graphs plotted here, which give the
993: functions along the $y$-axis, we find
994: of $||w^y||_\infty \approx 0.03$, while $||w^x||_\infty \approx 3
995: \times 10^{-3}$, and $||\phi -1||_\infty \approx 0.013$.
996: Because of the symmetry of
997: the configuration, $||w^z||_\infty$ is much smaller.
998: Analytically, $w^z = 0$ on the $y$-axis;
999: computationally we find $||w^z||_\infty \approx 5 \times 10^{-7}$.
1000: In fact we find in
1001: general that the radial component of $w^i$ is the dominant function
1002: in all directions, consistent with our boundary conditions, and
1003: consistent with the finding that solution of
1004: the constraints has small effect on the computed angular momentum.
1005: Of course the corrections $\phi$ and $w^i$ would be expected
1006: to be larger, for data describing holes closer together. We show below that
1007: this data setting method leads to generically smaller corrections than
1008: found in other methods, thus allowing closer control of the physical
1009: content of the data.
1010:
1011: \begin{figure}
1012: \begin{center}
1013: \includegraphics[width=3.0in, angle=270]{fig4.ps}
1014: \end{center}
1015: \caption{$\phi$ along the y-axis connecting two nonspinning holes with
1016: orbital angular momentum. The holes are boosted in the $\pm$
1017: x direction with $v=0.196$ and are separated by $10$~M. Note that $\phi$
1018: is very close to unity everywhere.
1019: }
1020: \label{fig:circular_orbit_nospin_1phi}
1021: \end{figure}
1022:
1023: \begin{figure}
1024: \begin{center}
1025: \includegraphics[width=3.0in, angle=270]{fig5.ps}
1026: \end{center}
1027: \caption{$w^{x}$ for the same configuration as in Figure
1028: \ref{fig:circular_orbit_nospin_1phi}.
1029: }
1030: \label{fig:circular_orbit_nospin_1wx}
1031: \end{figure}
1032:
1033:
1034: \begin{figure}
1035: \begin{center}
1036: \includegraphics[width=3.0in, angle=270]{fig6.ps}
1037: \end{center}
1038: \caption{$w^{y}$ for the same configuration as in
1039: Figure \ref{fig:circular_orbit_nospin_1phi}. $w^z$ is numerically
1040: zero as expected by symmetry.
1041: }
1042: \label{fig:circular_orbit_nospin_1wy}
1043: \end{figure}
1044:
1045: %----------------------------------------------------------------------
1046: %
1047: %
1048: %
1049: %----------------------------------------------------------------------
1050: \section{Binding Energy in Initial Data}
1051: \label{sec:binding_energy}
1052:
1053: As a first step towards understanding the physical content of initial
1054: data sets,
1055: we examine in this section the effect of the presence of a second hole
1056: on the horizon areas of a first hole and on global features such
1057: as the ADM integrals and the binding energy of the pair.
1058: This analysis is carried out for non-spinning holes to first order in the
1059: binding energy.
1060: A comparison to the Newtonian result indicates that the
1061: Kerr-Schild {\it background} superposition data contain the appropriate
1062: physical information at this level.
1063: We then consider possible spin-related phenomena, estimate
1064: their magnitude, and discuss their possible effect near the ISCO.
1065:
1066:
1067: %----------------------------------------------------------------------
1068: %
1069: %
1070: %
1071: %----------------------------------------------------------------------
1072: \subsection{Binding energy in Brill-Lindquist data}
1073: \label{sec:brill_lindquist}
1074:
1075:
1076: Before discussing the conformal Kerr-Schild data, we first consider
1077: Brill-Lindquist data for two non-moving Schwarzschild black holes~\cite{Brill}.
1078: These data are conformally flat, and $K_{ab} = 0$. The momentum
1079: constraints are trivially satisfied, and the Hamiltonian constraint is
1080: solved for a conformal factor: $\phi = 1 + m/(2r) + m'/(2r')$.
1081: Here the two mass parameters are $m$, and $m'$,
1082: and $r$ and $r'$ are the distances in the flat background
1083: from the holes $m$ and $m'$.
1084:
1085: We find that
1086: the apparent horizon areas in the solved data correspond to
1087: \be
1088: M_{\rm AH} + M_{\rm AH}' = m + m' + \frac{mm'}{\ell} +
1089: \rmO(\ell^{-2}).
1090: \ee
1091: The subscript
1092: ``$\rm AH$''
1093: indicates masses computed from apparent horizon areas, and the
1094: separation in the flat background space is $\ell$~\cite{cadez}.
1095: We assume that this mass (computed from {\it apparent} horizons)
1096: is close to the total intrinsic mass of the black holes
1097: (which is given by a knowledge of the spin---here zero---and the area of the
1098: {\it event} horizon).
1099: The
1100: binding energy, $\BE$, can be computed as the difference of the ADM mass
1101: observed at infinity and the sum of the horizon masses:
1102: \be
1103: \BE = M_{\ADM} - M_{\rm AH} - M'_{\rm AH}.
1104: \ee
1105: For Brill-Lindquist data $M_{\ADM} = m + m'$, so that
1106: %
1107: \be
1108: \BE = -\frac{Gmm'}{\ell} +\rmO(\ell^{-2})\, ,
1109: \label{eq:be_brill_lind}
1110: \ee
1111: which is the Newtonian result.
1112:
1113: %----------------------------------------------------------------------
1114: %
1115: %
1116: %
1117: %----------------------------------------------------------------------
1118: \subsection{Binding Energy in Superposed Kerr-Schild Data}
1119:
1120: We now calculate the binding energy in superposed Kerr-Schild data (set
1121: according to our conformal transverse-traceless approach) for a
1122: non-moving Schwarzschild black hole at the origin, and a second such
1123: hole at coordinate distance $\ell$ away. ($\ell$ is measured in the
1124: flat space associated with the data construction.) We compute the area
1125: of the hole at the origin to first order and find that the Newtonian
1126: binding energy already appears in the the background data prior
1127: to solving the constraints. Thus, we have an argument justifying the
1128: result noted at the end of \sref{sec:boundary}: solving the elliptic
1129: constraint equations leads to small corrections to the
1130: Kerr-Schild background data.
1131:
1132: Let both holes be placed on the $z$-axis; the first
1133: hole with mass parameter $m$ at the origin, and a second hole with
1134: mass parameter $m'$ at $z=\ell$. The holes are well separated, and
1135: we expand all quantities about the origin in powers
1136: of $\epsilon \equiv m'/\ell$ with
1137: $\epsilon \ll 1$. Using Schwarzschild coordinates labeled
1138: $(r, \theta, \phi)$ (cf.\ \eref{eq:1}--\eref{eq:6} for $a=0$),
1139: the background metric tensor is
1140: \beq
1141: \tilde g_{rr} &=& 1 + \frac{2m}{r} + 2\epsilon\cos^2\theta,\\
1142: \tilde g_{r\theta} &=& -2\epsilon r \sin\theta \cos\theta,\\
1143: \tilde g_{\theta \theta} &=& r^2 + 2\epsilon r^2 \sin^2\theta,\\
1144: \tilde g_{\phi \phi} &=& r^2\sin^2\theta,
1145: \eeq
1146: with all other components zero. The extrinsic curvature of the second
1147: hole, ${}_2 K_{ab}$, is of $\rmO(\epsilon^2)$ at the origin, and we have
1148: simply $\tilde K_{ab}={}_1 K_{ab}$.
1149: %\beq
1150: %\tilde K_{rho\rho} &=& -c_1 \frac{\rho+M}{\rho^3},\\
1151: %\tilde K_{\theta\theta} &=& c_1,\\
1152: %\tilde K_{\rho\rho} &=& c_1 \sin^2\theta,
1153: %\eeq
1154: %where
1155: %\be
1156: %c_1 \equiv 2M\sqrt{\frac{\rho}{\rho+2M}}.
1157: %\ee
1158: Similarly, the trace of the extrinsic curvature is $\tilde K = {}_1 K$.
1159: %\be
1160: %\tilde K = {}_1 K = c_1 \frac{(\rho+3M)}{\rho^2(\rho+2M)}.
1161: %\ee
1162: Finally, the non-zero components of $\tilde A_{ab}$ are
1163: \beq
1164: \tilde A_{r r} &=& -\frac{2c_2}{r^2}
1165: \lp 1 + 2\frac{m}{r} + 2\epsilon \cos^2\theta \rp,\\
1166: %
1167: \tilde A_{r \theta } &=& \epsilon \frac{c_2}{r}\sin\theta \cos\theta,\\
1168: %
1169: \tilde A_{\theta \theta } &=& c_2 \lp 1 + 2\epsilon \sin^2\theta \rp,\\
1170: %
1171: \tilde A_{\phi \phi} &=& c_2\sin^2\theta.
1172: \eeq
1173: where
1174: \be
1175: c_2 \equiv \frac{2M}{3} \sqrt{\frac{\rho}{\rho+2M}} \frac{(2r+3m)}{(r+2m)}.
1176: \ee
1177: While $\tilde K_{ab}$ is not a function of $\epsilon$, and hence
1178: contains no information about the second hole, perturbative quantities
1179: do appear in $\tilde A_{ab}$. This perturbation in $\tilde A_{ab}$ arises
1180: because we sum the mixed-index components of ${}_AA^c_b$, and because the
1181: full background metric,
1182: involving terms from each hole,
1183: is involved in the symmetrization in~\eref{eq:ks_super}.
1184:
1185:
1186: To calculate the binding energy we first find the apparent horizon area
1187: of the local hole. For a single Schwarzschild hole, the horizon
1188: is spherical and located at $\rho_{\HOR} = 2m$; the area of the horizon
1189: is $16\pi m^{2}$. The effect of the second hole is to distort the
1190: horizon along the $z$-axis connecting them, and we define a trial apparent
1191: horizon surface as $f=0$, where
1192: %
1193: \be
1194: f = \rho - 2m - \sum_{l} a_{l} P_{l} (\cos \theta)\,.
1195: \label{eq:f}
1196: \ee
1197: The expansion of $f$ in Legendre polynomials, $P_{l}$,
1198: expresses the distortion of the local horizon away from the zero-order
1199: spherical result.
1200: This expansion includes a term describing a constant
1201: ``radial"
1202: offset in the position of the apparent horizon, $a_{0}P_{0}$.
1203: This and the other terms defining the surface have the expected
1204: magnitude, %$a_l= \rmO \left(\frac{mm'}{\ell} \right)$.
1205: $a_l= \rmO (m\epsilon)$.
1206: We solve for the horizon by placing this expression for the
1207: surface into the apparent horizon equation
1208: %
1209: \begin{equation}
1210: \nabla_{i}s^{i} + A_{ab} s^{a}s^{b} - \frac{2}{3} K = 0,
1211: \label{eq:app_horizon}
1212: \end{equation}
1213: %
1214: where $s^{i}$ is the unit normal to the trial surface
1215: \be
1216: s_{i} = \frac{f_{,i}}{\sqrt{g^{ab}f_{,a}f_{,b}}}.
1217: \ee
1218:
1219: %\be
1220: % F_0(\rho) = \frac{2}{\rho} \frac{1-\frac{2m}{\rho}}{1+\frac{2m}{\rho}},.
1221: %\label{30}
1222: %\ee
1223:
1224: The apparent horizon equation is solved to first order, $\rmO(\epsilon)$.
1225: One must evaluate the equation at the new
1226: (perturbed) horizon location. Let $F$ represent the left-hand side of
1227: the apparent horizon equation (\eref{eq:app_horizon}), $\rho_0=2m$ is
1228: the horizon surface of the single, unperturbed hole, and
1229: $\rho_\HOR(\theta)$ is the new perturbed horizon. We expand $F$
1230: to first order as
1231: \be
1232: F(\rho_\HOR(\theta)) = F_0(\rho_0)
1233: + \frac{\partial F}{\partial \rho}\biggr|_{\rho_0} \sum a_{l}
1234: P_{l}\, = 0.
1235: \label{31}
1236: \ee
1237: Solving (\eref{eq:app_horizon}), the
1238: only nonzero coefficients in Eq.(\ref{eq:f})
1239: are $a_0=mm'/(3\ell),$ $a_2= -mm'/(2\ell)$.
1240: Integrating the determinant of the perturbed metric over the horizon
1241: surface, $\rho = 2m + \sum_{l} a_{l} P_{l} (\cos \theta)$, we find
1242: the area of the apparent horizon to be
1243: \be
1244: A_\HOR = 16\pi \lp m + \frac{mm'}{2\ell}\rp^{2}
1245: + \rmO\lp m^2(m'/\ell)^2\rp,
1246: \label{32}
1247: \ee
1248: corresponding to a horizon mass of $M_\HOR = m + mm'/(2\ell)$ to Newtonian
1249: order, ie to order $\rmO(\epsilon) = \rmO(\ell^{-1})$.
1250:
1251: In this nonmoving case the total ADM mass is just $M_\ADM = m + m'$.
1252: This leads to the Newtonian binding energy at this order
1253: \be
1254: \BE = -\frac{mm'}{\ell}.
1255: \label{be}
1256: \ee
1257:
1258: Because we work only to lowest order in $\epsilon$, \eref{be} had to result
1259: in an expression of $\rmO(m\epsilon)$, but it did not have to have
1260: a coefficient of unity.
1261: Both the conformally flat and conformally Kerr-Schild data
1262: contain the Newtonian binding energy.
1263: However, this result is obtained in the superposed {\it background}
1264: Kerr-Schild
1265: metric, while the Brill-Lindquist and \v Cade\v z data give
1266: the correct binding energy only after solving the elliptic constraints.
1267: This is consistent with the small corrections introduced by $\phi$ and
1268: $w^i$ ($\phi \sim 1$, $|w^i| \ll 1$) in the solved Kerr-Schild data
1269: (see \sref{sec:boundary}).
1270: This fact---that for a superposed Kerr-Schild background the solution of the
1271: full elliptic problem modifies the data (and the mass/angular momentum
1272: computations) only
1273: slightly---demonstrates how powerful this choice of data can be.
1274:
1275:
1276: Furthermore, the Newtonian form of the binding energy ($\epsilon \ll
1277: 1$) means the correct classical total energy is found for orbiting
1278: situations. If the holes have nonrelativistic motion, their individual
1279: masses are changed by order $\gamma \approx 1+ \rmO(v^2) = 1 + \rmO(\epsilon)$.
1280: The binding
1281: energy, which is already $\rmO(\epsilon )$ and is proportional to the product
1282: of the masses, is changed only at order $\rmO(\epsilon^2)$. The ADM
1283: mass, on the other hand, measures $\gamma m$, and $M_\ADM$ will be
1284: increased by $m v^2/2$ (an $\rmO(\epsilon)$ increase) for each hole,
1285: leading to the correct Newtonian energetics for the orbit.
1286:
1287: The apparent horizon is the only structure available to measure the
1288: intrinsic mass of a black hole. Complicating this issue is the intrinsic
1289: spin of the black hole; the relation is between horizon area and
1290: {\it irreducible} mass:
1291:
1292: \be
1293: A_{\rm H} = 16 \pi m_{irr}^2 = 8 \pi m \lp m + \sqrt{(m^2 -a^2)}\rp
1294: \label{eq:mirr}
1295: \ee
1296: %
1297: As \eref{eq:mirr} shows, the irreducible mass is a function of both the
1298: mass and the spin, and in general we cannot specify the spin of the
1299: black holes. For axisymmetric cases Ashtekar's isolated horizon
1300: paradigm~\cite{Dreyer}
1301: gives a way to measure the spin locally. We do not pursue the point here
1302: since we investigate generic and typically non-axisymmetric situations.
1303:
1304:
1305: %----------------------------------------------------------------------
1306: %
1307: %
1308: %
1309: %----------------------------------------------------------------------
1310: \subsection{Spin effects in Approximating Inspiral With Initial Data Sequences}
1311: \label{sec:spineffects}
1312:
1313: We have seen that the initial data contain the binding energy in a multiple
1314: black hole spacetime. This information can be used to deduce some
1315: characteristics of the orbital dynamics, particularly the radius of
1316: the circular orbit, $\ell$, and the orbital frequency, $\omega$.
1317: Given a sequence of initial data slices with
1318: decreasing separations, we determine $\BE$ for each slice.
1319: The circular orbit is found where
1320: \be
1321: \frac{\partial \BE}{\partial \ell}\biggr|_{J} = 0.
1322: \ee
1323: The separation at the ISCO orbit, $\ell_\ISCO$, lies at the boundary between
1324: binding energy curves which have a minimum, and those that do not. The
1325: curve for the ISCO has an inflection point:
1326: \be
1327: \frac{\partial^2 \BE}{\partial \ell^2}\biggr|_{J} = 0.
1328: \ee
1329: The
1330: angular frequency is given by
1331: \be
1332: \omega_\ISCO = \frac{\partial \BE}{\partial J} .
1333: \label{omega}
1334: \ee
1335:
1336:
1337: The attempt to model dynamical inspiral seems secure for large
1338: separation ($\ell>15 m$), though
1339: surprises appear even when the holes are very well separated.
1340: For instance, \eref{32} above shows that
1341: compared to the bare parameter values, the
1342: mass increase is equal for the two holes in a
1343: dataset. Thus the smaller hole is proportionately more strongly affected
1344: than the larger one is.
1345:
1346: The physically measurable quantity in question is the frequency (at
1347: infinity) \eref{omega} associated with the last orbit prior to the plunge, the
1348: ISCO. This may be impossible to determine by the initial data set method.
1349:
1350: To begin with, isolated black holes form a 2-parameter set (depending
1351: on the mass parameter $m$, and the angular momentum parameter,
1352: $j=ma$). For isolated black holes without charge
1353: the parameters $\{m, j\} $ uniquely
1354: specify the hole. They are equal, respectively, to the
1355: physical mass and angular momentum. Every method of constructing
1356: multiple black hole data assigns parameter values ${}_Bm$ and ${}_Bj$ to
1357: each constituent ${}_B(hole)$ in the data set.
1358:
1359: There is substantial ambiguity involving spin and mass in setting the
1360: black hole data. One must consider the evolutionary
1361: development of the black hole area and spin. This is a real physical
1362: phenomenon which contradicts at some
1363: level the usual assumption of invariant mass and spin. A related
1364: concern arises because it is only the {\it total} ADM angular momentum that
1365: is accessible in the data, whereas one connects to particle motion via
1366: the {\it orbital} angular momentum.
1367:
1368: Consider the behavior of the individual black hole spin and mass
1369: in an inspiral. For widely separated holes, because
1370: the spin effects fall off faster with distance than the dominant
1371: mass effects do, we expect the spin to be approximately conserved in an
1372: inspiral. Therefore it should also be constant across the initial data sets
1373: representing a given sequence of orbits. But when the holes approach closely,
1374: the correct choice of spin parameter becomes problematic also.
1375:
1376: Newtonian arguments demonstrate some of the possible spin effects. In every
1377: case they are {\it a priori} small until the orbits approach very
1378: closely. However, at estimates for the ISCO, the effects begin to be
1379: large, and result in ambiguities in setting the data (see Price and
1380: Whelan~\cite{P+W}).
1381: We will consider these effects in decreasing order of their magnitude.
1382:
1383: For two holes, each of mass $m$ in Newtonian orbit with a total separation of
1384: $\ell$, the orbital frequency is
1385:
1386: \be
1387: m \omega = \sqrt 2 (m/\ell)^{(3/2)}.
1388: \label{35c}
1389: \ee
1390:
1391: From recent work by Pfeiffer et al.~\cite{Pfeiffer}, the estimated
1392: ISCO frequency is of order $m\Omega=0.085$, corresponding to $\ell \approx
1393: 6.5\, m$ in this Newtonian approach.
1394:
1395:
1396: To compare this frequency, \eref{35c},
1397: to an intrinsic frequency in the problem, we take the lowest
1398: (quadrupole) quasi-normal mode of the final merged black hole (of mass
1399: $2m$) which has frequency $2m\omega_0 \approx 0.37$; the quadrupole
1400: distortion is excited at twice the orbital frequency. (We are using the
1401: values for a Schwarzschild black hole in this qualitative analysis.) The
1402: driving frequency equals the quasi-normal mode frequency when
1403: $\ell \approx 4\, m$, as might be expected.
1404:
1405: To consider effects linked to the orbital motion on the initial
1406: configurations, we can first treat the effect of imposing corotation.
1407: While we show below that corotation is not physically enforced except
1408: for very close orbits, it is a fact that certain formulations, for
1409: instance versions of the ``thin sandwich" with a helical Killing vector,
1410: require corotation in their treatment. For any particular
1411: initial orbit, corotation is certainly a possible situation.
1412:
1413: In corotation, then, with \eref{35c}, for each hole:
1414: %
1415: \be
1416: J = m a = I\omega = 4 m^2 (m \omega).
1417: \label{35dd}
1418: \ee
1419: %
1420: The result for the moment of inertia $I=4m^3$ is the Schwarzschild
1421: value~\cite{MTW,membrane}. Thus
1422: %
1423: \be
1424: a = 4m \sqrt{2} (m/\ell)^{(3/2)}.
1425: \label{35e}
1426: \ee
1427: %
1428: Assume $a/m \ll 1$, and compute the area of this black hole~\cite{MTW}:
1429: %
1430: \be
1431: A= 8 \pi m(m+\sqrt{m^2-a^2})\\
1432: \approx 16\pi m^2 (1- (a/m)^2/4).
1433: \label{35ff}
1434: \ee
1435: %
1436: The horizon mass computed from this area is
1437: \be
1438: \sqrt{A/(16\pi)} \approx m(1-4(m/\ell)^3).
1439: \label{35f}
1440: \ee
1441:
1442: At our estimate of the ISCO orbit, $\ell_\ISCO\approx 6\, m$, this effect is
1443: of order of 10\% of the Newtonian binding energy, distinctly enough to affect
1444: the location of the ISCO.(At $\ell_\ISCO\approx 6\, m$, $a/m \approx 0.3$
1445: for corotation).
1446:
1447: Two more physical effects are not typically considered in setting data.
1448: They are {\it frame dragging}, and {\it tidal torquing}. Within our
1449: Newtonian approximations, we will find that these effects are small, but
1450: not zero as the orbits approach the ISCO.
1451: In full nonlinear gravity these effects could be substantial precisely
1452: at the estimated ISCO.
1453:
1454: The frame dragging is the largest dynamical
1455: effect. The orbiting binary possesses a net angular momentum. For a
1456: rotating mass (here the complete binary system) the frame dragging
1457: angular rate is estimated as the rotation rate times the
1458: gravitational potential at the measurement point~\cite{MTW}. Hence
1459: %
1460: \be
1461: m\Omega_{\rm{drag}}= m\omega\lp \frac{2m}{\ell}\rp
1462: \approx \lp\frac{m}{\ell}\rp^{\frac{5}{2}}
1463: \approx \frac{a}{4m}.
1464: \label{35gg}
1465: \ee
1466: %
1467: This is $a/m$ of order 1\% at $\ell=10m$; of order 4\% at $\ell=6m$.
1468:
1469:
1470: The tidal torquing and dissipative heating of the black holes can be
1471: similarly estimated.
1472: As the two holes spiral together, the
1473: tidal distortion from each hole on the other will have a frequency
1474: which is below, but approaching the quasi-normal frequency. Just as for
1475: tidal effects in the solar system, there will be lag in the phase angle
1476: of the distortion, which we can determine because the lowest
1477: quasi-normal mode is a dissipative oscillator, driven through the tidal
1478: effects at twice the orbital frequency:
1479: %
1480: \be
1481: \ddot q +2 \gamma \dot q + \omega_0^2 q= F(\omega).
1482: \label{35d}
1483: \ee
1484: %
1485: Here $m^2q$ is the quadrupole moment of the distorted black hole. The
1486: parameter $\gamma$ is (for a Schwarzschild hole of mass $2m$)) about
1487: $2m \gamma = 0.089$. In \eref{35d} the driving acceleration $F(\omega)$
1488: is identified with the tidal distortion acceleration. We evaluate it at zero
1489: frequency:
1490: %
1491: \beq
1492: q &=& F(\omega=0)/\omega_0^2\\
1493: &\approx & \frac{m}{\ell^3}.
1494: \label{36}
1495: \eeq
1496: %
1497: The lagging phase, for driving frequency $2\omega \ll \omega_0$, is easily computed
1498: to be
1499: %
1500: \beq
1501: \phi &\approx& 4 \gamma \frac{\omega}{\omega_0^2}\\
1502: &=& 4\lp\frac{\gamma}{\omega_0}\rp \lp\frac{\omega}{\omega_0}\rp
1503: \label{37}
1504: \eeq
1505: %
1506: This lagging tidal distortion will produce a tidal torque on the black
1507: hole, which we can approximate using a combination of Newtonian and
1508: black hole ideas. The most substantial approximation is that the torque
1509: arises from a redistribution of the mass in the ``target" black hole, of
1510: amount $\Delta m = m m^2q = m (m/\ell)^3$. This mass has separation
1511: $\approx 4m$. Thus the torque on the hole is
1512: %
1513: \begin{eqnarray}
1514: \tau &=& \sin \phi \times (\mbox{lever arm}) \times \Delta F \\ \nonumber
1515: &=& \sin \phi \times (4m) \times (\Delta m 2m^2 / \ell^3) \\ \nonumber
1516: &=& 8 \sin \phi m(m/\ell)^6\\ \nonumber
1517: &\approx & 32 (\gamma/\omega_0)(\omega/\omega_0)m(m/\ell)^6\\ \nonumber
1518: &\approx& 60 m(m/\ell)^{15/2}.
1519: \label{38}
1520: \end{eqnarray}
1521:
1522: What is most important is the effect of this torque on the angular
1523: momentum of the hole over the period of time it takes the orbit to
1524: shrink from a very large radius. To accomplish this, we use the inspiral
1525: rate (calculated under the assumption of weak gravitational radiation
1526: from the orbit; see~\cite{MTW}):
1527: %
1528: \be
1529: \frac{d\ell}{dt} = -\frac{128}{5} \lp \frac{m}{\ell}\rp^3
1530: \label{39}
1531: \ee
1532: %
1533: Thus
1534: %
1535: \begin{eqnarray}
1536: \frac{dJ}{d\ell} &=& \tau \frac{dt}{d\ell} \\
1537: &=& -\frac{5\tau}{128} \lp \frac{m}{\ell}\rp^{-3} \\
1538: &\approx& -2 m \lp \frac{m}{\ell}\rp^{9/2},
1539: \label{40}
1540: \end{eqnarray}
1541: %
1542: and
1543: %
1544: \be
1545: J(\ell) \approx m^2 \left(\frac{m}{\ell}\right)^{7/2};
1546: \label{40x}
1547: \ee
1548: %
1549: assuming that there is minimal mass increase from the associated
1550: heating (which we discuss just below), this identifies the induced spin
1551: parameter $a= m(m/\ell)^{7/2}$ for an inspiral from infinity.
1552:
1553:
1554: The estimate $a= m(m/\ell)^{7/2}$ for an inspiral from infinity assumes the
1555: mass of the hole has not changed significantly in the inspiral. By
1556: considering the detailed behavior of the shear induced in the horizon
1557: by the tidal perturbation, the growth in the black hole mass can be
1558: estimated~\cite{membrane} as
1559: %
1560: \begin{eqnarray}
1561: \frac{dm}{dt} &=& \omega \frac{dJ}{dt},
1562: \label{41}
1563: \end{eqnarray}
1564: %
1565: leading to a behavior
1566: %
1567: \be
1568: \Delta m(\ell) \approx 5 m \lp \frac{m}{\ell}\rp^{5};
1569: \label{42}
1570: \ee
1571: %
1572: Consequently, the change in mass can be ignored until the holes are quite
1573: close. However, the point is that these Newtonian estimates lead to
1574: possible strong effects just where they become unreliable, and just
1575: where they would affect the ISCO.
1576:
1577: These results are consistent with similar ones of Price and Whelan~\cite{P+W},
1578: who estimated tidal torquing using a derivation due to
1579: Teukolsky~\cite{TeukThesis}. That derivation assumes the quadrupole
1580: moment in the holes arises from their Kerr character, which predicts
1581: specific values for the quadrupole moment,
1582: as a function of angular momentum parameter $a$.
1583:
1584: Finally we consider an effect on binding energy shown by Wald and also by Dain.
1585: Wald directly computes the force for stationary sources with
1586: arbitrarily oriented spins. He considered a small black hole as a
1587: perturbation in the field of a large hole. The result found~\cite{WaldPRD} was
1588:
1589: \be
1590: \BE = - \frac{mm'}{\ell} - \lp \frac{ \vec{S} \cdot \vec{S'} -
1591: 3(\vec{S} \cdot \hat{n})(\vec{S'} \cdot \hat{n})}{\ell^3} \rp.
1592: \label{BEWald}
1593: \ee
1594:
1595: \noindent Here, $\vec{S}$, $\vec{S}'$ are the spin vectors of the
1596: sources and $\hat{n}$ is the unit vector connecting the two sources.
1597: Dain~\cite{Dain}, using a definition of intrinsic mass that differs
1598: from ours, finds binding energy which agrees with Wald's \eref{BEWald}
1599: at $\rmO(\ell^{-3})$. This is discussed further in Section
1600: \ref{sec:physresults}.
1601:
1602: %----------------------------------------------------------------------
1603: %
1604: %
1605: %
1606: %----------------------------------------------------------------------
1607: \section{Numerical Results}
1608:
1609: We now turn to computational solutions of the constraint equations
1610: to generate physical data using the superposed Kerr-Schild data.
1611: We first discuss the computational code and tests, as well as
1612: some of the limitations of the code. Finally, we consider
1613: physical conclusions that can be drawn from the results.
1614:
1615: %----------------------------------------------------------------------
1616: %
1617: %
1618: %
1619: %----------------------------------------------------------------------
1620: \subsection{Code Performance}
1621:
1622:
1623: The constraint equations are solved (\eref{ell_eqs}) with an accelerated
1624: SOR solver~\cite{NumRecepies}. The solution is iterated until the $L_{2}$
1625: norms of the residuals
1626: of the fields are less than $10^{-10}$, far below truncation error.
1627: Discrete derivatives are approximated with second order, centered derivatives.
1628: We are limited to fairly small domains, e.g., $x^i \in [-12 m, 12m]$
1629: for a typical $m/8$ resolution using $193^3$ points.
1630:
1631: To verify the solution of the discrete equations, we have examined the
1632: code's convergence in some detail. The constraints have known analytical
1633: solutions---they should be zero---which allows us to determine the code's
1634: convergence using a solution calculated at two different resolutions.
1635: Let $S_1$ be a solution calculated with resolution $h_1$, and
1636: $S_2$ be a solution calculated with $h_2$, then the convergence
1637: factor $c_{12}$ is
1638: \begin{eqnarray}
1639: c_{12} = \frac{\log\left(\frac{||S_{1}||}
1640: {||S_{2}||}\right)}{\log\left(\frac{h_{1}}{h_{2}}\right)}
1641: \label{eq:convCalc}
1642: \end{eqnarray}
1643:
1644: We constructed a conformal background spacetime with two $m=1$ non-spinning
1645: black holes separated by $6m$ on the $y$-axis.
1646: The elliptic equations were then solved on grids with resolutions of
1647: $m/6$, $m/8$, $m/10$ and $m/12$.
1648: Tables \ref{tab:TableHam}--\ref{tab:TableMomX} show the convergence factors
1649: as a function of resolution for the Hamiltonian constraint and
1650: the $x$-component of the momentum constraint, ${\cal C}^x$. The
1651: convergence for ${\cal C}^y$ is nearly identical to ${\cal C}^x$, and
1652: as the $y$-axis is an axis of symmetry, ${\cal C}^z$ is identical to
1653: ${\cal C}^x$.
1654: Figures \ref{fig:ham2}--\ref{fig:momz2} show the convergence behavior of the
1655: constraints along coordinate lines. The constraints
1656: calculated at lower resolutions are rescaled to the highest resolution by
1657: the ratio of resolutions squared. We see second order for all components
1658: with the exception of the points nearest to the inner boundary.
1659:
1660:
1661: \begin{table}
1662: \begin{tabular}{|l|ccc|}\hline
1663: \multicolumn{4}{|c|}{Convergence $(c_{ab})$}\\
1664: ~~ & ~~$a=m/6$ & ~~$a=m/8$ & ~~$a=m/10$ \\ \hline
1665: $b=m/8$ & 1.70 & ~~ & ~~ \\
1666: $b=m/10$ & 1.77 & 1.86 & ~~ \\
1667: $b=m/12$ & 1.79 & 1.86 & 1.85 \\ \hline
1668: \end{tabular}
1669: \vspace{.5in}
1670: \label{tab:TableHam}
1671: \caption{Convergence data for the Hamiltonian constraint,
1672: ${\cal C}^0$, for a solution with two $m=1$, non-spinning holes
1673: at $x^i=(0,\pm 3m,0)$ in the conformal background, and outer boundaries
1674: at $x^i=\pm 6m$. The solution was calculated at resolutions $m/6$,
1675: $m/8$, $m/10$, and $m/12$. The $L_2$ norms of ${\cal C}^0$ were calculated
1676: over the entire volume of the domain
1677: using a mask of radius $1m$ around each hole, while the computational mask
1678: has a radius of approximately $0.75~m$. This larger mask was used to
1679: compensate for the slight difference of physical location of the mask
1680: at different resolutions. The norms are as follows:
1681: $||{\cal C}^0(m/6)||_2 = 0.00389054$,
1682: $||{\cal C}^0(m/8)||_2 = 0.00238321$,
1683: $||{\cal C}^0(m/10)||_2 = 0.00157387$ and
1684: $||{\cal C}^0(m/12)||_2 = 0.00112328$.
1685: }
1686: \end{table}
1687:
1688: \begin{table}
1689: \begin{tabular}{|l|ccc|}\hline
1690: \multicolumn{4}{|c|}{Convergence $(c_{ab})$}\\
1691: ~~ & ~~$a=m/6$ & ~~$a=m/8$ & ~~$a=m/10$ \\ \hline
1692: $b=m/8$ & 1.93 & ~~ & ~~ \\
1693: $b=m/10$ & 1.99 & 2.06 & ~~ \\
1694: $b=m/12$ & 1.99 & 2.03 & 1.99 \\ \hline
1695: \end{tabular}
1696: \vspace{.5in}
1697: \label{tab:TableMomX}
1698: \caption{Convergence data for the $x$-component of the
1699: momentum constraint, for the same configuration as Table~\ref{tab:TableHam}.
1700: The norms of ${\cal C}^x$ are as follows:
1701: $||{\cal C}^x(m/6)||_2 = 0.00541231$,
1702: $||{\cal C}^x(m/8)||_2 = 0.00310937$,
1703: $||{\cal C}^x(m/10)||_2 = 0.00196156$ and
1704: $||{\cal C}^x(m/12)||_2 = 0.00136514$.
1705: Convergence factors were also calculated for ${\cal C}^y$ and ${\cal C}^z$,
1706: and found to be essentially identical to the data shown here, and thus
1707: are not given separately.
1708: }
1709: \end{table}
1710:
1711: \begin{figure}
1712: \begin{center}
1713: \includegraphics[width=3.in,angle=270]{fig7.ps}
1714: \end{center}
1715: \caption{The Hamiltonian constraint (units $m^{-2}$) along $y$-axis after
1716: solving the elliptic equations for 4 different levels of resolution. The
1717: constraints are rescaled by the ratio of the resolutions squared,
1718: showing second
1719: order convergence. The two non-spinning, instantaneously stationary
1720: holes of $m=1$ are positioned at $\pm3$ on the $y$-axis.}
1721: \label{fig:ham2}
1722: \end{figure}
1723:
1724: \begin{figure}
1725: \begin{center}
1726: \includegraphics[width=3.in, angle=270]{fig8.ps}
1727: \end{center}
1728: \caption{$y$-component of momentum constraint (units $m^{-2}$) along the $y$-axis after
1729: solving the elliptic equations for 4 different levels of resolution,
1730: showing second order convergence. The background physical
1731: situation is the same as in
1732: Figure (\ref{fig:ham2}). The other momentum constraint components evaluated
1733: on this axis are zero by symmetry, both analytically, and computationally
1734: ($\rmO(10^{-12})$).}
1735: \label{fig:momy2}
1736: \end{figure}
1737:
1738:
1739: \begin{figure}
1740: \begin{center}
1741: \includegraphics[width=3.in, angle=270]{fig9.ps}
1742: \end{center}
1743: \caption{$z$-component of momentum constraint (units $m^{-2}$) along the $z$-axis after
1744: solving the elliptic equations for 4 different levels of resolution,
1745: showing second order convergence.
1746: Other components of the momentum constraint evaluated along this line
1747: are zero by symmetry, both
1748: analytically and computationally ($\rmO(10^{-12})$).
1749: The background physical situation
1750: is the same as in
1751: Figure (\ref{fig:ham2}). The behavior of the $x$-momentum constraint
1752: along the $x$-axis is identical to this Figure, as required by the
1753: symmetry of the problem.}
1754: \label{fig:momz2}
1755: \end{figure}
1756:
1757:
1758: %\begin{table}
1759: %\begin{tabular}{cc}\hline\hline
1760: %\multicolumn{2}{c}{Momentum$_{y}$:}\\\hline\hline
1761: %Resolution & L2 norm \\
1762: %m/6 & 0.00551741 \\
1763: %m/8 & 0.00315638 \\
1764: %m/10 & 0.00198946 \\
1765: %m/12 & 0.00138186
1766: %\end{tabular}
1767:
1768: %\begin{tabular}{lccc}\hline
1769: %\multicolumn{4}{c}{Convergence $(c_{ab})$}\\
1770: %~~ & a=6 & a=8 & a=10 \\ \hline
1771: %b=8 & 1.94131925291174 & ~~ & ~~ \\
1772: %b=10 & 1.99685617693102 & 2.06845572486942 & ~~ \\
1773: %b=12 & 1.9973797316871 & 2.03715527437679 & 1.99884661903467 \\ \hline
1774: %\end{tabular}
1775: %\vspace{.5in}
1776: %\label{tab:TableMomY}
1777: %\end{table}
1778:
1779: %\begin{table}
1780: %\begin{tabular}{cc}\hline\hline
1781: %\multicolumn{2}{c}{Momentum$_{z}$:}\\\hline\hline
1782: %Resolution & L2 norm \\
1783: %m/6 & 0.00541231 \\
1784: %m/8 & 0.00310937 \\
1785: %m/10 & 0.00196156 \\
1786: %m/12 & 0.00136514
1787: %\end{tabular}
1788:
1789: %\begin{tabular}{lccc}\hline
1790: %\multicolumn{4}{c}{Convergence $(c_{ab})$}\\
1791: %~~ & a=6 & a=8 & a=10 \\ \hline
1792: %b=8 & 1.9266263300294 & ~~ & ~~ \\
1793: %b=10 & 1.98685396059259 & 2.0645008810491 & ~~ \\
1794: %b=12 & 1.98719556241069 & 2.0301701166937 & 1.9881526569248 \\ \hline
1795: %\end{tabular}
1796: %\vspace{.5in}
1797: %\label{tab:TableMomZ}
1798: %\end{table}
1799:
1800:
1801:
1802:
1803: %------------------------------------------------
1804: \begin{table}
1805: \begin{tabular}{|c|c|}
1806: \hline \hline
1807: Domain & $M_{ADM}$ \\
1808: \hline
1809: $\pm$ 8 m & 1.942 m \\
1810: $\pm$ 10 m & 1.964 m \\
1811: $\pm$ 11 m & 1.974 m \\
1812: $\pm$ 12 m & 1.980 m \\
1813: \hline \hline
1814: \end{tabular}
1815: \caption{Total ADM Mass for two instantaneously stationary, non-spinning
1816: holes separated by $6 m$ on a grid of discretization $\Delta x = m/8$ for
1817: four different domain sizes.}
1818: \label{tab:table1}
1819: \end{table}
1820: %------------------------------------------------
1821:
1822: Solutions of elliptic equations are well-known to be dependent on all
1823: boundary data. The outer boundary is an artificial boundary, as the
1824: the physical spacetime is unbounded. Boundary data for this outer
1825: boundary are derived from the asymptotic behavior of a single Kerr
1826: black hole. On very large domains these conditions should closely
1827: approximate the expected field behavior, but on small domains
1828: these boundary data may only crudely approximate the real solution.
1829: This error in the boundary data then contaminates the entire solution,
1830: as expected for elliptic solutions. Additional error arises in the
1831: calculation of the ADM quantities, as spacetime near the outer boundary
1832: does not approach asymptotic flatness.
1833: As an indication of the error associated
1834: with the artificial outer boundaries, we calculated solutions with the
1835: same physical parameters on grids of differing sizes. The boundary
1836: effects in the $M_\ADM$ are given in \tref{tab:table1}, and
1837: \fref{fig:phi_zero} shows a contour plot of $\phi$ for equal mass,
1838: nonspinning, instantaneously stationary black holes with the outer
1839: boundaries at $x^i = \pm 12m$. As a further demonstration of
1840: boundary effects in our solutions, \fref{fig:phi_zero10} shows
1841: $\phi$ for a configuration examined by Pfeiffer {\it et al.}~\cite{Pfeiffer2}.
1842: Their solution, shown in Fig.~8 of~\cite{Pfeiffer2}, was computed on a
1843: much larger domain via a spectral method~\cite{Kidder}.
1844: Thus, while we achieve reasonable results,
1845: it is important to remember that the boundary effects may be significant.
1846: Moreover, we have only considered the effect of outer boundaries, while
1847: errors arising from the approximate inner boundary condition have
1848: not been examined.
1849:
1850: \begin{figure}
1851: \begin{center}
1852: \includegraphics[width=3.0in]{fig10.ps}
1853: \end{center}
1854: \caption{
1855: Contour plot of $\phi$ for two instantaneously stationary, non-spinning
1856: holes of mass parameter $m = 1$.
1857: The holes are separated by $6~m$ along the y-axis. The bold circles indicate
1858: the apparent horizons.}
1859: \label{fig:phi_zero}
1860: \end{figure}
1861:
1862:
1863: \begin{figure}
1864: \begin{center}
1865: \includegraphics[width=3.0in]{fig11.ps}
1866: \end{center}
1867: \caption{
1868: Contour plot representing the same configuration as \fref{fig:phi_zero}
1869: but with the holes separated by
1870: $10$~m along the y-axis. Compare to Figure 8 in ref.~\cite{Pfeiffer2}}.
1871: \label{fig:phi_zero10}
1872: \end{figure}
1873:
1874: \begin{figure}
1875: \begin{center}
1876: \includegraphics[width=3.in,angle=270]{fig12.ps}
1877: \end{center}
1878: \caption{The total ADM energy for two momentarily stationary non-spinning
1879: black holes separated by $6 m$ at various resolutions.
1880: The results exhibit second order convergence.}
1881: \label{fig:2bh_adm_mass}
1882: \end{figure}
1883:
1884:
1885: \begin{figure}
1886: \begin{center}
1887: \includegraphics[width=3.in,angle=270]{fig13.ps}
1888: \end{center}
1889: \caption{The total ADM angular momentum for two non-spinning holes
1890: boosted in the $\pm$
1891: x direction with $v=0.3162$ and separated by $6 m$ at various
1892: resolutions (background angular momentum $\tilde J^{ADM}_{12} = 2.0 m^2$).}
1893: \label{fig:2bh_adm_angMom}
1894: \end{figure}
1895:
1896:
1897:
1898: \begin{figure}
1899: \begin{center}
1900: \includegraphics[width=3.in]{fig14.ps}
1901: \end{center}
1902: \caption{Conformal factor $\phi$ for two instantaneously stationary holes
1903: separated by $6m$
1904: with spin parameter $a=0.5$. The spins are parallel and pointed out of
1905: the page. Compare to \fref{fig:phi_zero}. Also notice the boundary
1906: effect on the outermost contour, labeled $0.999$. }
1907: \label{fig:phi_spinaligned}
1908: \end{figure}
1909:
1910:
1911: \begin{figure}
1912: \begin{center}
1913: \includegraphics[width=3.in]{fig15.ps}
1914: \end{center}
1915: \caption{Conformal factor $\phi$ for the same configuration
1916: as \Fref{fig:phi_spinaligned} except the spins are anti-parallel:
1917: the the spin of the hole at $(0,-3,0)$ points into the page.}
1918: \label{fig:phi_spinantialigned}
1919: \end{figure}
1920:
1921:
1922:
1923: \begin{figure}
1924: \begin{center}
1925: \includegraphics[width=2.5in]{fig16.ps}
1926: \end{center}
1927: \caption{$\phi$ for a grazing collision between
1928: two equal mass, non-spinning holes. The holes are centered at $y = \pm 1 m$
1929: and boosted toward each
1930: other with $v_x = \mp 0.5c$, respectively. }
1931: \label{fig:grazing}
1932: \end{figure}
1933:
1934:
1935: \begin{figure}
1936: \begin{center}
1937: \includegraphics[width=2.5in]{fig17.ps}
1938: \end{center}
1939: \caption{$\phi$ for two non-spinning holes boosted
1940: perpendicularly to their separation. The holes are separated by $10 m$
1941: and boosted
1942: with $v_x = \pm 0.196$, giving the system a background
1943: angular momentum of $\tilde J^{ADM}_{12} = 2.0 m^2$.
1944: The calculated $J^{ADM}_{12} = 1.91 m^2$ and
1945: $M_{ADM} = 1.970 m^2$. The Newtonian data correspond to an elliptic orbit
1946: at apastron. }
1947: \label{fig:circular}
1948: \end{figure}
1949:
1950:
1951: Other derived quantities also show convergence: \fref{fig:2bh_adm_mass}
1952: shows the ADM mass $M_{\ADM}$ for two nonspinning black holes at $6m$
1953: separation, and different resolutions.
1954: The fit is
1955: \be
1956: M_{ADM} = \lp 1.941 + 0.067 \lp \frac{\Delta x}{m}\rp - 0.422
1957: \lp \frac{\Delta x}{m} \rp^2 \rp m.
1958: \label{eq:42S}
1959: \ee
1960: %
1961: showing good second order convergence. The angular momentum calculation
1962: is less robust, but exhibits approximately first order convergence.
1963: The fit to Figure \ref{fig:2bh_adm_angMom}, which shows $J^{ADM}$ for
1964: two nonspinning holes with orbital angular momentum, gives:
1965:
1966: \be
1967: J^{ADM}_{12} = \lp 1.837 -
1968: 0.121 \lp \frac{\Delta x}{m}\rp +
1969: 0.237 \lp \frac{\Delta x}{m} \rp^2 \rp \epsilon_{12}m^2.
1970: \label{eq:42T}
1971: \ee
1972: %
1973: Compare this to the angular momentum computed
1974: for the background: $\tilde J^{ADM}_{12} = 2.0 m^2$.
1975:
1976:
1977:
1978: %----------------------------------------------------------------------
1979: %
1980: %
1981: %
1982: %----------------------------------------------------------------------
1983: \subsection{Physics Results}
1984: \label{sec:physresults}
1985:
1986: The small computational domain does not negate the
1987: utility of these solutions as initial data for the time-dependent
1988: Einstein equations. For instance, Figures \ref{fig:grazing} and
1989: \ref{fig:circular} show data for grazing and elliptical orbits.
1990: They are currently being incorporated into the Texas binary black hole
1991: evolution code. While the small domains do mean that
1992: that our data do not represent the best asymptotically flat results
1993: available from this method, we can still verify some of the qualitative
1994: analytical predictions of the previous section.
1995: In particular, Figures \ref{fig:phi_zero}, \ref{fig:phi_spinaligned},
1996: and \ref{fig:phi_spinantialigned}
1997: show the conformal factor $\phi$
1998: for holes instantaneously at rest at a separation of $6m$.
1999: In Figure \ref{fig:phi_zero} they are
2000: non-spinning; in Figures
2001: \ref{fig:phi_spinaligned} and
2002: \ref{fig:phi_spinantialigned}, each has Kerr parameter $a=0.5m$.
2003: In one case (\fref{fig:phi_spinaligned}) the spins are aligned;
2004: in the other (\fref{fig:phi_spinantialigned}) they are antialigned.
2005: \tref{tab:table2} gives
2006: the values of the apparent horizon area of each hole, the ADM mass, and the
2007: binding energy fraction for these configurations.
2008: The binding energy is consistent with the
2009: analytic estimates of Wald~\cite{WaldPRD} in Section \ref{sec:spineffects}.
2010:
2011: Wald's computation of the binding energy for spinning holes,
2012: \eref{BEWald}, gives for parallel or antiparallel spins orthogonal to
2013: the separation (as in our computational models):
2014:
2015: \be
2016: \BE = - \lp \frac{mm'}{\ell} + \frac{ \vec{S} \cdot \vec{S'}}{\ell^3} \rp
2017: \label{BEwald}
2018: \ee
2019: %
2020: with oppositely directed spins showing {\it less} binding energy.
2021: For our $S = 0.5~m^2$, $\ell = 6~m$ configuration, this is a
2022: change of order $\rmO(2 \times 10^{-3})$ between the parallel and the
2023: antiparallel cases. For the spinning cases we compute a
2024: change in binding energy between parallel and antiparallel
2025: spins of roughly half that,
2026: with the correct sign. This rough
2027: correspondence to the analytic result is suggestive. However, the nonspinning
2028: case deviates from the expectation that its binding energy
2029: should be between that
2030: of the spinning cases. Based on the scatter in the binding energies
2031: shown here, we estimate that we have achieved about $3 \%$ accuracy in the
2032: binding energy. With the accuracy of our solution and the size of our domain,
2033: we are unable to present a clearer dependence of binding energy on spin.
2034:
2035: \begin{table}
2036: \begin{tabular}{l || l | l | l | l}
2037: \hline \hline
2038: & $A^{\dag}_{AH}$ & $M_{ADM}$ & $M^{\dag}_{AH}$ & Binding Energy \\
2039: \hline \hline
2040: Parallel Spin & 53.20 & 1.973 & 1.065 & -0.157 = -0.147 $\times M_{AH}$ \\
2041: Antiparallel Spin & 53.17 & 1.974 & 1.065 & -0.156 = -0.146 $\times M_{AH}$ \\
2042: Zero Spin & 57.10 & 1.980 & 1.066 & -0.151 = -0.142 $\times M_{AH}$ \\
2043: \hline \hline
2044: \end{tabular}
2045: \caption{$M_{ADM}$, $A_{AH}$ and associated quantities
2046: calculated for two holes with $m = 1.0$ on a grid $(24~m)^3$ with
2047: resolution $\Delta x = m/8$. \\
2048: ${}^\dag$: Quantity for a single hole.\\}
2049: \label{tab:table2}
2050: \end{table}
2051:
2052:
2053: %----------------------------------------------------------------------
2054: %
2055: %
2056: %
2057: %----------------------------------------------------------------------
2058: \section{Outlook}
2059:
2060:
2061:
2062: To an extent, the difficulty in setting data will become less relevant,
2063: as good evolutions are eventually achieved. Then data can be set for
2064: initial configurations with very large separation, and the subsequent
2065: evolution will tell us the future dynamics. In the shorter term, the
2066: iBBH program of Thorne and collaborators~\cite{Brady}
2067: will give us an indication of
2068: the evolution of the black hole parameters in the inspiral, and will
2069: allow a closer identification of the corresponding initial data
2070: sequence. To point out a couple of additional physical effects,
2071: note that, besides the historical component associated
2072: with a lagging tidal distortion, there is the familiar fact that most
2073: data setting methods are incapable of accounting for the previously
2074: emitted gravitational radiation. One can then expect that data
2075: describing hyperbolic encounters will be more accurate than data sets
2076: describing circular motion. This is because, in hyperbolic encounters, which are
2077: set as distant initial configurations, the radiation is more planar,
2078: and confined to near each hole.
2079: The radiation should then both be better defined, and should
2080: have less effect on the subsequent evolution than in the more distorted
2081: orbiting data set. In any case, the
2082: understanding of these problems is extremely significant in
2083: understanding the physical content of the configurations we must solve
2084: to provide waveforms for the new generation of gravitational wave
2085: detectors.
2086: For more accurate computational results, we are undertaking a both
2087: multigrid approach~\cite{Klasky}, and a spectral approach~\cite{Pfeiffer2, GGB1}
2088: and expect to have extended results soon comparable to those
2089: of~\cite{Pfeiffer2,Kidder}. An ultimate goal of such a solver is to be
2090: able to carry
2091: out elliptic solutions at every integration time step, to enable fully
2092: constrained evolutions.
2093:
2094:
2095:
2096:
2097: %----------------------------------------------------------------------
2098: %
2099: %
2100: %
2101: %----------------------------------------------------------------------
2102: %\subsection{Apparent Horizons}
2103:
2104:
2105: %----------------------------------------------------------------------
2106: %
2107: %
2108: %
2109: %----------------------------------------------------------------------
2110: %\section{Conclusion}
2111:
2112:
2113:
2114: \section*{Acknowledgments}
2115:
2116: RM thanks A. Ashtekar, A. \v Cade\v z, G, Cook, P. Laguna, H.
2117: Pfeiffer, and D. Shoemaker for insightful comments.
2118: %We especially thank
2119: %P. Marronetti for extensive discussions on elliptic solver technique,
2120: %and for a critique of an earlier version of this work.
2121: This work is
2122: supported in part by NSF grants PHY~9800722, PHY~9800725, and
2123: PHY~0102204. Computations were performed at the NSF supercomputer
2124: center NCSA and the University of Texas AHPCC.
2125:
2126:
2127: \newpage
2128:
2129: \begin{thebibliography}{99}
2130:
2131: \bibitem{Cook} G.~B.~Cook, {\it Phys.\ Rev.} {\bf D50}, 5025 (1994).
2132:
2133: \bibitem{Pfeiffer} H.~P.~Pfeiffer, S.~A.~Teukolsky and G.~B.~Cook
2134: {\it Phys.\ Rev.} {\bf D62}, 104018 (2000).
2135:
2136: \bibitem{Baumgarte}T.~Baumgarte {\it Phys. Rev.} {\bf D62}, 024018 (2000).
2137:
2138: \bibitem{GGB1} E.~Gourgoulhon, P.~Grandclement and S.~Bonazzola,
2139: {\it Phys.\ Rev.} {\bf D65}, 044020 (2002).
2140:
2141: \bibitem{GGB2}
2142: P.~Grandclement, E.~Gourgoulhon and S.~Bonazzola,
2143: {\it Phys.\ Rev.} {\bf D65}, 044021 (2002), and references therein.
2144:
2145: \bibitem{GGB3} E.~Gourgoulhon, P.~Grandclement and S.~Bonazzola,
2146: {\it Int.\ J.\ Mod.\ Phys.} {\bf A17} 2689--94 (2002).
2147:
2148: \bibitem{Cook2} G.~B.~Cook {\it Phys.\ Rev.} {\bf D65} 084003 (2002).
2149:
2150: \bibitem{Pfeiffer2} H.~P.~Pfeiffer, G.~B.~Cook, and S.~A.~Teukolsky
2151: {\it Phys.\ Rev.} {\bf D66} 024047 (2002).
2152:
2153:
2154:
2155: \bibitem{CookReview} G.~B.~Cook, {\it Living Rev.\ Rel.} {\bf 3}, 5 (2000).
2156:
2157:
2158:
2159:
2160: \bibitem{Peters} P.~Peters and J.~Matthews,
2161: {\it Phys. Rev.} {\bf 131},435 (1963).
2162:
2163:
2164: \bibitem{Will} T.~Mora and C.~Will, {\it Phys.\ Rev.} {\bf D66}, 101501 (2002).
2165:
2166:
2167: \bibitem{Shoemaker}D.~Shoemaker, M.~F.~Huq and R.~A.~Matzner
2168: {\it Phys. Rev.} {\bf D62}, 124005 (2000).
2169:
2170: \bibitem{Huq}M.~F.~Huq, M.~Choptuik and R.~A.~Matzner,
2171: {\it Phys. Rev.} {\bf D66}, 084024 (2002). [arXiv:gr-qc/0002076].
2172:
2173:
2174: \bibitem{Dreyer} O.~Dreyer, B.~Krishnan, E.~Schnetter, and D.~Shoemaker,
2175: {\it Phys.\ Rev.} {\bf D67}, 024018 (2003), and references therein.
2176:
2177: \bibitem{Caveny} Scott~A.~Caveny, PhD Dissertation,
2178: University of Texas at Austin (2002).
2179:
2180: \bibitem{Caveny1}Scott~A.~Caveny, Matthew Anderson and Richard~A.~Matzner
2181: ``Tracking Black Holes in Numerical Relativity", submitted to {\it Phys. Rev.}
2182: (2003). [arXiv:gr-qc/0303099].
2183:
2184: \bibitem{ADM} R.~Arnowitt, S.~Deser, and C.~Misner in Witten,
2185: {\it Gravitation,
2186: an Introduction to Current Research} (Wiley, New York 1962).
2187:
2188: \bibitem{YP} J.~York and T.~\ Piran
2189: ``The Initial Value Problem and Beyond'',
2190: \textit{Spacetime and Geometry: The Alfred Schild
2191: Lectures}, R.~Matzner and L.~Shepley Eds.
2192: University of Texas Press, Austin, Texas. (1982);
2193: G.~Cook, ``Initial Data for the Two-Body Problem
2194: of General Relativity'', Ph.D. Dissertation, The University
2195: of North Carolina at Chapel Hill (1990).
2196:
2197:
2198: \bibitem{MY} N.~{\'O}.~Murchadha and J.~W.~York, Jr.,
2199: {\it Phys.\ Rev.} {\bf D10}, 428 (1974);
2200: N.~{\'O}.~Murchadha and J.~W.~York, Jr.,
2201: {\it Phys.\ Rev.} {\bf D10}, 437 (1974);
2202: N.~{\'O}.~Murchadha and J.~W.~York, Jr.,
2203: {\it Gen.\ Relativ.\ Gravit.} {\bf 7} 257 (1976).
2204:
2205: \bibitem{York} J.~W.~York, Jr., {\it Phys.\ Rev.\ Lett.} {\bf 82},
2206: 1350 (1999).
2207:
2208:
2209:
2210: \bibitem{Mathews} J.~R.~Wilson and G.~J.~Mathews,
2211: {\it Phys.~Rev.~Lett.} {\bf 75}, 4161 (1995);
2212: J.~R.~Wilson and G.~J.~Mathews, and P.~Marronetti,
2213: {\it Phys.~Rev.} {\bf D54}, 1317 (1996)
2214:
2215:
2216:
2217:
2218:
2219: \bibitem{Bowen+York} J.~Bowen and J.~W.~York,
2220: {\it Phys. Rev.} {\bf D21}, 2047 (1980).
2221:
2222:
2223: \bibitem{Garat}
2224: A.~Garat and R.~H.~Price,
2225: %``Nonexistence of conformally flat slices of the Kerr spacetime,''
2226: {\it Phys.\ Rev.} {\bf D61}, 124011 (2000) [arXiv:gr-qc/0002013].
2227:
2228: \bibitem{Jansen}
2229: N.~Jansen, P.~Diener, A.~Khokhlov and I.~Novikov,
2230: ``Local and global properties of conformally flat
2231: initial data for black hole collisions,'' {\it Class. Quant. Grav.}
2232: {\bf 20} 51 (2003) [arXiv:gr-qc/0103109].
2233:
2234: \bibitem{Matzner:1999pt}
2235: R.~Matzner, M.~F.~Huq and D.~Shoemaker,
2236: {\it Phys.\ Rev.} {\bf D59}, 024015 (1999)[arXiv:gr-qc/9805023].
2237:
2238:
2239:
2240:
2241: \bibitem{KerrSchild} R.~Kerr and A.~Schild, ``Some Algebraically Degenerate
2242: Solutions of Einstein's Gravitational Field Equations,'' in
2243: {\it Applications of Nonlinear Partial Differential Equations in
2244: Mathematical Physics}, Proc. of Symposia B Applied Math., Vol XVII (1965);
2245: ``A New Class of Solutions of the Einstein Field Equations'',
2246: {\it Atti del Congresso Sulla Relitivita Generale: Problemi Dell'Energia
2247: E Onde Gravitazionala} G. Barbera, Ed. (1965).
2248:
2249: \bibitem{MTW} C.~W.~Misner, K.~S.~Thorne, and J.~A.~Wheeler, {\it Gravitation}
2250: (W.H. Freeman, New York, 1970).
2251:
2252: \bibitem{Marronetti:2000rw}
2253: P.~Marronetti, M.~Huq, P.~Laguna, L.~Lehner, R.~Matzner and
2254: D.~Shoemaker, {\it Phys.\ Rev.} {\bf D62}, 024017 (2000) [gr-qc/0001077].
2255:
2256: \bibitem{Marronetti}P.~Marronetti and R.~A.~Matzner
2257: {\it Phys. Rev. Lett.} {\bf 85} 5500 (2000).
2258:
2259: \bibitem{Wald} R.~ Wald, {\it General Relativity}
2260: (University of Chicago Press, Chicago, 1984).
2261:
2262:
2263: \bibitem{note4} For consistency of notation we
2264: use the superscript ``ADM" to decorate the angular momentum
2265: $J^{\ADM}_{ab}$. We are not certain where the formula (\ref{eq:adm_ang_mom})
2266: was first written. It does {\it not} appear in~\cite{ADM}.
2267: Expressions equivalent to it appear in~\cite{Weinberg}
2268: and in~\cite{MTW}, and were promulgated in notes by Misner (unpublished) in the
2269: mid-60s, arising from the long history of conservation law/pseudo-tensor
2270: studies in General Relativity. The form here was taken from~\cite{Wald}; see
2271: also the very thorough development in~\cite{Brown}.
2272:
2273: Further, objections have
2274: been raised to our use of the symbol $M_{\ADM}$, rather than $E_{\ADM}$ in
2275: \eref{eq:adm_mass}. For a moving source, \eref{eq:adm_mass} would yield
2276: $\gamma$ times the result for the same source not moving (a calculation
2277: facilitated by the Kerr-Schild structure); see the discussion around
2278: \eref{eq:boosted} below. We urge the reader to
2279: carefully follow our notation.
2280:
2281: \bibitem{Weinberg} S.~Weinberg, {\it Gravitation and Cosmology}
2282: (Wiley and Sons, New York, 1972)
2283:
2284: \bibitem{Brown} J.~D.~Brown and J.~W.~York, {\it Phys. Rev.} {\bf D47},
2285: 1407 (1993).
2286:
2287: \bibitem{note1} Spatial components of angular momentum (e.g. spin)
2288: perpendicular to the motion transform with one power of $\gamma$ and
2289: stay perpendicular to the motion. The orbital angular momentum ${\bf L}
2290: = {\bf r}\times {\bf p}$ also contains one power of $\gamma$ in ${\bf
2291: p}$. Hence ${\bf J}$ contains one power of $\gamma$. See Landau and
2292: Lifshitz, {\it Classical Theory of Fields} revised second edition,
2293: (Pergamon Press, Oxford, 1962), p. 46.
2294:
2295:
2296: \bibitem{Brill} D.~Brill and R.~W.~Lindquist
2297: {\it Phys. Rev.} {\bf 131} 471 (1963).
2298:
2299: \bibitem{cadez}
2300: This was apparently first noticed as a computational result; see
2301: A. \v Cade\v z, {\it Ann. Phys. (NY)} {\bf 83} 449 (1974). \v Cade\v z
2302: considered both the Brill-Lindquist~\cite {Brill} and
2303: Misner~\cite{Misner} data. For Brill-Lindquist data the apparent
2304: horizon areas are easy to compute to $\rmO(m m'/\ell)$, because the lowest
2305: order effect of the distant hole on the local one is a constant, isotropic
2306: addition to the local value of the conformal factor.
2307:
2308: \bibitem{Misner} C.~Misner, {\it Phys.\ Rev.} {\bf 118}, 1110 (1960).
2309:
2310:
2311: \bibitem{P+W} R.~H.~Price and J.~T.~Whelan,
2312: {\it Phys.\ Rev.\ Lett.} {\bf 87}, 231101 (2001)
2313: [arXiv:gr-qc/0107029].
2314:
2315: \bibitem{membrane}{\it Black Holes, the Membrane Paradigm}, K.~S.~Thorne,
2316: R.~H.~Price, D~.A.~Macdonald, editors. (Yale
2317: University Press, New Haven, 1986), Section VIII B 1.
2318:
2319: \bibitem{TeukThesis}S.~A.~Teukolsky, PhD Dissertation, California
2320: Institute of Technology, (1973).
2321:
2322: \bibitem{WaldPRD} R.~Wald, {\it Phys.\ Rev.} {\bf D6}, 406 (1972).
2323:
2324:
2325: \bibitem{Dain} S.~Dain {\it Phys. Rev.}{\bf D66} 084019 (2002).
2326:
2327:
2328: \bibitem{NumRecepies} W.~H.~Press, S.~A.~Teukolsky, W.~T.~Vetterling,
2329: and B.~P.~Flannery,
2330: {\it Numerical Recipies in Fortran, Second Edition} (Cambridge University
2331: Press, Cambridge, 1992).
2332:
2333:
2334: \bibitem{Kidder} H.~Pfeiffer L.~Kidder, M.~Scheel and S.~A.~Teukolsky
2335: [arXiv:gr-qc 0202096].
2336:
2337: \bibitem{Brady}P.~Brady, J.~Creighton and K.~S.~Thorne,
2338: {\it Phys. Rev.} {\bf D58}, 061501 (1998).
2339:
2340: \bibitem{Klasky} S.~Klasky, PhD Dissertation, University of Texas at Austin,
2341: (1966).
2342:
2343:
2344: \end{thebibliography}
2345: \end{document}
2346: \end
2347:
2348: