1: % $Id: ID.tex,v 1.31 2002/06/07 12:02:01 harald Exp $
2: \documentclass[aps,prd,showpacs,twocolumn, groupedaddress,nofootinbib]{revtex4}
3:
4: \usepackage{amsmath, amsfonts, graphicx}
5:
6: \newcommand{\ve}[1]{{\bf #1}}
7: \newcommand{\Long}{\mathbb{L}}
8: \newcommand{\LapLong}{\Delta_\Long}
9: \newcommand{\tildeLapLong}{\tilde\Delta_\Long}
10: \newcommand{\TF}{\mbox{TF}}
11:
12:
13: \begin{document}
14: \title{Comparing initial-data sets for binary black holes}
15:
16: \author{Harald P. Pfeiffer}
17: \affiliation{Department of Physics, Cornell University, Ithaca, New York\ \ 14853}
18: \author{Gregory B. Cook}
19: \affiliation{Department of Physics, Wake Forest University, Winston-Salem,
20: North Carolina\ \ 27109}
21: \author{Saul A. Teukolsky}
22: \affiliation{Columbia Astrophysics Laboratory, Columbia University, New York, New York 10027}
23: \altaffiliation[Permanent address: ]{Department of Physics, Cornell University, Ithaca, New York\ \ 14853}
24:
25: \date{\today}
26:
27: \begin{abstract}
28: We compare the results of constructing binary black hole initial
29: data with three different decompositions of the constraint equations
30: of general relativity. For each decomposition we compute the
31: initial data using a superposition of two Kerr-Schild black holes to
32: fix the freely specifiable data. We find that these initial-data sets
33: differ significantly, with the ADM energy varying by as much as 5\%
34: of the total mass. We find that all initial-data sets currently
35: used for evolutions might contain unphysical gravitational radiation
36: of the order of several percent of the total mass. This is
37: comparable to the amount of gravitational-wave energy observed
38: during the evolved collision. More astrophysically realistic
39: initial data will require more careful choices of the freely
40: specifiable data and boundary conditions for both the metric and
41: extrinsic curvature. However, we find that the choice of extrinsic
42: curvature affects the resulting data sets more strongly than the
43: choice of conformal metric.
44: \end{abstract}
45: \pacs{04.25.Dm, 04.70.Bw} \maketitle
46:
47:
48: \section{Introduction}
49:
50: Numerical evolutions of black holes have been improved slowly but
51: steadily over the last few years and now first attempts are being made
52: to extract physical information from these evolutions. Most notably
53: one wants to predict the gravitational radiation emitted during black
54: hole coalescence \cite{Baker-Bruegmann-etal:2001,Alcubierre:2001,
55: Baker-Campanelli-etal:astroph0202469}.
56:
57: The quality of the initial data will be crucial to the success of the
58: predictions of the gravitational wave forms. Unphysical gravitational
59: radiation present in the initial data will contribute to the
60: gravitational waves computed in an evolution and might overwhelm the
61: true gravitational wave signature of the physical process under
62: consideration. Therefore an important question is how to control the
63: gravitational wave content of initial-data sets, and how to specify
64: {\em astrophysically} relevant initial data with the appropriate
65: gravitational wave content, for e.g.\ two black holes orbiting each
66: other. Unfortunately, assessing and controlling the gravitational wave
67: content of initial-data sets is not well understood at all.
68:
69: The mere {\em construction} of an initial-data set alone is
70: fairly involved, since every initial-data set must satisfy a rather
71: complicated set of four partial differential equations, the so-called
72: constraint equations of general relativity. The question of how to
73: solve these equations, and how to specify initial data representing
74: binary black holes in particular, has received considerable attention.
75:
76: We consider in this paper three different approaches that transform
77: the constraint equations into elliptic equations: The
78: {\em conformal transverse-traceless (TT) decomposition}\cite{York:1979},
79: the {\em physical TT decomposition} \cite{ Murchadha-York:1974a,
80: Murchadha-York:1974b, Murchadha-York:1976} and the {\em conformal
81: thin sandwich decomposition}\cite{York:1999}. These decompositions
82: split the variables on the initial-data surface into various pieces in
83: such a way that the constraint equations determine some of the pieces while
84: not restricting the others. After these freely
85: specifiable pieces are chosen, the constraint equations are solved and the
86: results are combined with the freely specifiable pieces to yield a
87: valid initial-data set.
88:
89: Any reasonable choice for the freely specifiable pieces will lead to a
90: valid initial-data set. Furthermore, any one of these decompositions
91: can generate any desired initial-data set, given the {\em correct}
92: choices of the freely specifiable pieces. However, it is not clear
93: {\em what} choices of freely specifiable pieces lead to initial-data
94: sets with the desired properties.
95:
96: The decompositions we consider here lead to four coupled nonlinear
97: elliptic partial differential equations. Since such equations are difficult to
98: solve, the early approach to constructing initial data was pragmatic:
99: One used the conformal TT decomposition with additional restrictions
100: on the freely specifiable pieces, most notably conformal flatness and
101: maximal slicing. These assumptions decouple the constraints
102: and allow for analytical solutions to the momentum
103: constraints, the so-called {\em Bowen-York extrinsic
104: curvature}\cite{Bowen:1979,Bowen-York:1980,Kulkarni-Shepley-York:1983}.
105: All that remains is to solve a single elliptic
106: equation, the Hamiltonian constraint. This approach has been used in
107: several variations\cite{Thornburg:1987,Cook-Choptuik-etal:1993,
108: Brandt-Brugmann:1997}.
109:
110: However, these numerical simplifications come at a cost.
111: The freely specifiable pieces have been restricted to a small subset of
112: all possible choices. One therefore can generate only a subset of
113: all possible initial-data sets, one that might not contain the desired
114: astrophysically relevant initial-data sets.
115:
116: Over the last few years there have been additional developments:
117: Post-Newtonian results have indicated that binary black hole metrics
118: are not conformally flat\cite{Rieth:1997,
119: Damour-Jaranowski-Schaefer:2000b}.
120: With certain restrictions on the slicing, it has also been shown that a
121: single stationary spinning black hole cannot be represented with a
122: conformally flat spatial metric
123: \cite{Monroe:1976,Garat-Price:2000}. In \cite{Pfeiffer-Teukolsky-Cook:2000},
124: it was shown that conformally flat initial data sets for spinning
125: binary black holes contain an unphysical contamination. Moreover,
126: computations in spherical symmetry\cite{Lousto-Price:1998} indicated
127: that initial-data sets depend strongly on the choice of the extrinsic
128: curvature and that the use of the Bowen-York extrinsic curvature might
129: be problematic.
130:
131: Therefore it is necessary to move beyond conformally flat initial data
132: and to explore different choices for the extrinsic curvature. Matzner
133: et al\cite{Matzner-Huq-Shoemaker:1999} proposed a non-flat conformal
134: metric based on the superposition of two Kerr-Schild metrics; a
135: solution based on this proposal was obtained in
136: \cite{Marronetti-Matzner:2000}. This work demonstrated the existence
137: of solutions to the 3D set of equations, but did not examine the data
138: sets in any detail.
139: Refs.~\cite{Gourgoulhon-Grandclement-Bonazzola:2001a,Grandclement-Gourgoulhon-Bonazzola:2001b} obtained solutions to a
140: similar set of equations during the computation of quasi-circular
141: orbits of binary black holes. However, these works assumed conformal
142: flatness.
143:
144: In this paper we present a code capable of solving the three
145: above-mentioned decompositions of the constraint equations for
146: arbitrary choices of the freely specifiable pieces. This code is
147: based on spectral methods which have been used successfully for several
148: astrophysical problems (see
149: e.g. \cite{Bonazzola-Gourgoulhon-etal:1993,
150: Bonazzola-Gourgoulhon-Marck:1998,
151: Kidder-Scheel-etal:2000,
152: Gourgoulhon-Grandclement-etal:2001,
153: Kidder-Scheel-Teukolsky:2001,
154: Ansorg-Kleinwaechter-Meinel:2002,
155: Grandclement-Gourgoulhon-Bonazzola:2001b}). Our code is described in
156: detail in a separate paper\cite{Pfeiffer-Kidder-etal:2002}.
157:
158: We compute solutions of the different decompositions for the non-flat
159: conformal metric proposed in Ref.~\cite{Matzner-Huq-Shoemaker:1999}.
160: Each decomposition has certain choices for the freely specifiable
161: pieces and boundary conditions that seem ``natural'' and which we use
162: in our solutions. We compare the computed initial-data sets with each
163: other and with the ``standard'' conformally-flat solution using the
164: Bowen-York extrinsic curvature. Our major results confirm that
165: \begin{enumerate}
166: \item\label{res:1}the different decompositions generate different
167: physical initial-data sets for seemingly similar choices for the
168: freely specifiable pieces.
169: \item the choice of extrinsic curvature is critical.
170: \end{enumerate}
171: The first result is certainly not unexpected, but each of these
172: factors can cause relative differences of several per cent in
173: gauge-invariant quantities like the ADM-energy.
174:
175: We also find that the conformal TT/physical TT decompositions generate
176: initial-data sets with ADM-energies $2-3$\% higher than data sets of
177: the conformal thin sandwich decomposition. We demonstrate that this
178: higher ADM-energy is related to the choice of the freely specifiable part
179: of the extrinsic curvature. In addition, we find that the solutions
180: depend significantly on the boundary conditions used.
181:
182: The paper is organized as follows. In the next section we describe the
183: three decompositions. Section \ref{sec:FreePieces} explains how we
184: choose the freely specifiable data within each decomposition. In
185: section \ref{sec:Implementation} we describe and test our elliptic
186: solver. Section \ref{sec:Results} presents our results, which we
187: discuss in section \ref{sec:Discussion}.
188:
189:
190: \section{Decompositions of Einstein's equations and the constraint equations}
191:
192: \subsection{3+1 Decomposition}
193:
194: In this paper we use the standard 3+1 decomposition of Einstein's
195: equations. We foliate the spacetime with $t=\mbox{const}$
196: hypersurfaces and write the four-dimensional metric as
197: \begin{equation}
198: ^{(4)}ds^2=-N^2\,dt^2+\gamma_{ij}(dx^i+N^idt)(dx^j+N^jdt),
199: \end{equation}
200: where $\gamma_{ij}$ represents the induced 3-metric on the
201: hypersurfaces, and $N$ and $N^i$ represent the lapse function and the
202: shift vector, respectively. We define the extrinsic curvature $K_{ij}$
203: on the slice by
204: \begin{equation}\label{eq:K}
205: {\bf K}=-\frac{1}{2}\,{\perp}\:{\cal L}_n{^{(4)}{\bf g}}
206: \end{equation}
207: where ${}^{(4)}{\bf g}$ is the space-time metric, $n$ the unit
208: normal to the hypersurface, and $\perp$ denotes the projection operator
209: into the $t=\mbox{const}$ slice.
210: Einstein's
211: equations divide into constraint equations, which constrain the data
212: $(\gamma_{ij}, K^{ij})$ on each hypersurface, and into evolution
213: equations, which determine how the data $(\gamma_{ij}, K^{ij})$ evolve
214: from one hypersurface to the next.
215: The constraint equations are
216: \begin{align}
217: \label{eq:HamiltonianConstraintPhysical}
218: R+K^2-K_{ij}K^{ij}&=16\pi G\rho\\
219: \label{eq:MomentumConstraint}
220: \nabla_j\left(K^{ij}-\gamma^{ij}K\right)&=8\pi G j^i.
221: \end{align}
222: Eq.~(\ref{eq:HamiltonianConstraintPhysical}) is called the {\em
223: Hamiltonian constraint}, and Eq.~(\ref{eq:MomentumConstraint}) is
224: referred to as the {\em momentum constraint}. $K=\gamma_{ij}K^{ij}$
225: is the trace of the extrinsic curvature, $\nabla$ and $R$ denote the
226: three dimensional covariant derivative operator and the Ricci scalar
227: compatible with $\gamma_{ij}$. $\rho$ and $j^i$ are the energy and
228: momentum density, respectively. Both vanish for the vacuum spacetimes
229: considered here.
230:
231: The evolution equation for $\gamma_{ij}$ is
232: \begin{equation}
233: \label{eq:Evolution-gamma}
234: \partial_t\gamma_{ij}=-2N K_{ij}+\nabla_iN_j+\nabla_jN_i,
235: \end{equation}
236: which follows from Eq.~(\ref{eq:K}). There is a similar albeit
237: longer equation for $\partial_tK_{ij}$ which we will not need in this
238: paper. The choices of $N$ and $N^i$ are arbitrary. One can in
239: principle use any lapse and shift in the evolution off the
240: initial-data surface, although some choices of lapse and shift are better
241: suited to numerical implementation than others.
242:
243: Later in this paper we will often refer to the trace-free piece of
244: Eq.~(\ref{eq:Evolution-gamma}). Denote the tracefree piece of a tensor
245: by $\TF(.)$, and define $\gamma\equiv\det \gamma_{ij}$. From Eq.~(\ref{eq:Evolution-gamma}) and the fact that
246: $\delta\ln\gamma=\gamma^{kl}\delta\gamma_{kl}$, it follows that
247: \begin{equation}\label{eq:Evolution-gamma-TF}
248: \TF(\partial_t\gamma_{ij})
249: =\gamma^{1/3}\partial_t\left(\gamma^{-1/3}\gamma_{ij}\right)
250: =-2N A_{ij}+(\Long N)_{ij}.
251: \end{equation}
252: Here $A_{ij}=K_{ij}-\frac{1}{3}\gamma_{ij}K$ denotes the trace-free
253: extrinsic curvature, and
254: \begin{equation}
255: \label{eq:DefinitionLong}
256: (\Long N)^{ij}\equiv
257: \nabla^iN^j+\nabla^jN^i-\frac{2}{3}\gamma^{ij}\nabla_kN^k.
258: \end{equation}
259: $\Long$ always acts on a vector, so the 'N' in $(\Long N)^{ij}$ denotes
260: the shift vector $N^i$ and not the lapse $N$.
261:
262: \subsection{Decomposition of the constraint equations}
263: \label{sec:Decomposition}
264:
265: Equations~(\ref{eq:HamiltonianConstraintPhysical}) and
266: (\ref{eq:MomentumConstraint}) constrain four degrees of freedom of the
267: 12 quantities $(\gamma_{ij}, K^{ij})$. However, it is not immediately
268: clear which pieces of $\gamma_{ij}$ and $K^{ij}$ are constrained and
269: which pieces can be chosen at will. Several decompositions have been
270: developed to divide the 12 degrees of freedom into freely specifiable
271: and constrained pieces. We will now review some properties of
272: the three decompositions we consider in this paper.
273:
274: All three decompositions follow the York-Lichnerowicz approach and
275: use a conformal transformation on the
276: physical 3-metric $\gamma_{ij}$,
277: \begin{equation}
278: \label{eq:ConformalMetric}
279: \gamma_{ij}=\psi^4\tilde\gamma_{ij}.
280: \end{equation}
281: $\psi$ is called the {\em conformal factor}, $\tilde\gamma_{ij}$ the
282: {\em background metric} or {\em conformal metric}. We will denote all
283: conformal quantities with a tilde. In particular, $\tilde\nabla$ is
284: the covariant derivative operator associated with $\tilde\gamma_{ij}$,
285: and $\tilde R_{ij}$ and $\tilde R$ are the Ricci tensor and Ricci
286: scalar of $\tilde\gamma_{ij}$.
287:
288: The extrinsic curvature is split into its trace and trace-free part,
289: \begin{equation}
290: \label{eq:SplitK_Aij}
291: K^{ij}=A^{ij}+\frac{1}{3}\gamma^{ij}K.
292: \end{equation}
293: The three decompositions of the constraint equations we discuss in
294: this paper differ in how $A^{ij}$ is decomposed. For each
295: decomposition, we discuss next the relevant equations, and describe
296: how we choose the quantities one has to specify before solving the
297: equations. We use the conventions of \cite{Cook:2000}.
298:
299:
300: \subsubsection{Conformal TT Decomposition}
301:
302: In this decomposition one first conformally transforms the traceless
303: extrinsic curvature,
304: \begin{equation}
305: \label{eq:ConfTT-AijConfAij}
306: A^{ij}=\psi^{-10}\tilde A^{ij},
307: \end{equation}
308: and then applies a TT decomposition with respect to the background
309: metric $\tilde\gamma_{ij}$:
310: \begin{equation}
311: \label{eq:ConfTT-Aij}
312: \tilde A^{ij}=\tilde A_{TT}^{ij}+(\tilde\Long X)^{ij}.
313: \end{equation}
314: The operator $\tilde\Long$ is defined by Eq.~(\ref{eq:DefinitionLong}) but
315: using the conformal metric $\tilde\gamma_{ij}$ and derivatives
316: associated with $\tilde\gamma_{ij}$. $\tilde A^{ij}_{TT}$ is
317: transverse with respect to the conformal metric, $\tilde\nabla_j\tilde
318: A^{ij}_{TT}=0$, and is traceless.
319:
320: Substituting Eqs.~(\ref{eq:ConfTT-AijConfAij}) and (\ref{eq:ConfTT-Aij})
321: into the momentum constraint (\ref{eq:MomentumConstraint}), one finds
322: that it reduces to an elliptic equation for $X^i$, whereas $\tilde
323: A^{ij}_{TT}$ is unconstrained.
324:
325: In order to specify the transverse-traceless tensor $\tilde
326: A^{ij}_{TT}$ one usually has to {\em construct} it from a general
327: symmetric trace-free tensor $\tilde M^{ij}$ by subtracting the
328: longitudinal piece. As described in \cite{Cook:2000} one can
329: incorporate the construction of $\tilde A^{ij}_{TT}$ from $\tilde
330: M^{ij}$ into the momentum constraint, arriving at the following
331: equations:
332:
333: \begin{gather}
334: \label{eq:ConfTT-1}
335: \tilde\nabla^2\psi-\frac{1}{8}\psi\tilde R-\frac{1}{12}\psi^5K^2
336: +\frac{1}{8}\psi^{-7}\tilde A_{ij}\tilde A^{ij}=-2\pi G\psi^5\rho,\\
337: \label{eq:ConfTT-2}
338: \tildeLapLong V^i-\frac{2}{3}\psi^6\tilde\nabla^iK
339: +\tilde\nabla_j\tilde M^{ij}=8\pi G\psi^{10}j^i,
340: \end{gather}
341: where $\tilde A^{ij}$ and the operator $\tildeLapLong$ are defined by
342: \begin{equation}
343: \label{eq:ConfTT-Aij_tilde}
344: \tilde A^{ij}=(\tilde\Long V)^{ij}+\tilde M^{ij}
345: \end{equation}
346: and
347: \begin{equation}
348: \label{eq:DefinitionLapLong}
349: \tildeLapLong V^i\equiv\tilde\nabla_j(\tilde\Long V)^{ij}.
350: \end{equation}
351:
352: After solving these equations for $\psi$ and $V^i$, one obtains the
353: physical metric $\gamma_{ij}$ from (\ref{eq:ConformalMetric}) and the
354: extrinsic curvature from
355: \begin{equation}
356: \label{eq:ConfTT-3}
357: K^{ij}=\psi^{-10}\tilde A^{ij}+\frac{1}{3}\psi^{-4}\tilde\gamma^{ij}K.
358: \end{equation}
359: We will refer to Eqs.~(\ref{eq:ConfTT-1}) and (\ref{eq:ConfTT-2})
360: together with (\ref{eq:ConfTT-Aij_tilde}), (\ref{eq:ConfTT-3}) and
361: (\ref{eq:ConformalMetric}) as the {\em conformal TT equations}. In
362: these equations we are free to specify the background metric
363: $\tilde\gamma_{ij}$, the trace of the extrinsic curvature $K$, and a
364: symmetric traceless tensor $\tilde M^{ij}$. The solution $V^i$ will
365: contain a contribution that removes the longitudinal piece from
366: $\tilde M^{ij}$ and the piece that solves the momentum constraint if
367: $\tilde M^{ij}$ were transverse-traceless.
368:
369:
370: This decomposition has been the most important in the past, since if
371: one chooses a constant $K$ and if one considers vacuum spacetimes then
372: the momentum constraint (\ref{eq:ConfTT-2}) decouples from the
373: Hamiltonian constraint (\ref{eq:ConfTT-1}). Moreover, if one assumes
374: conformal flatness and $\tilde M^{ij}=0$, it is possible to write down
375: analytic solutions to Eq.~(\ref{eq:ConfTT-2}), the so-called
376: Bowen-York extrinsic curvature. In that case one has to deal with
377: only one elliptic equation for $\psi$. The Bowen-York extrinsic
378: curvature can represent multiple black holes with arbitrary momenta
379: and spins. One can fix boundary conditions for $\psi$ by requiring
380: that the initial-data slice be inversion symmetric at both
381: throats\cite{Misner:1963,Cook:1991}. In that case one has to modify the
382: extrinsic curvature using a method of images. We will
383: include initial-data sets obtained with this approach below, where we
384: will refer to them as {\em inversion symmetric} initial data.
385:
386: Reasonable choices for the freely specifiable pieces $\tilde\gamma_{ij}$, $K$,
387: $\tilde M^{ij}$ will lead to an initial-data set $(\gamma_{ij},
388: K^{ij})$ that satisfies the constraint equations. How should we choose
389: all these functions in order to obtain a desired physical
390: configuration, say a binary black hole with given linear momenta and
391: spins for the individual holes? We can gain insight into this question
392: by considering how the conformal TT decompositions can recover a known
393: solution.
394:
395: Suppose we have a known solution $(\gamma_{0\,ij}, K^{ij}_0)$
396: of the constraint equations. Denote the trace and trace-free parts of
397: this extrinsic curvature by $K_0$ and $A^{ij}_0$, respectively. If we
398: set
399: \begin{equation}
400: \tilde\gamma_{ij}=\gamma_{0\,ij},\quad K=K_0,
401: \quad\tilde M^{ij}=A^{ij}_0
402: \end{equation}
403: then
404: \begin{equation}
405: \psi=1,\quad V^i=0
406: \end{equation}
407: trivially solve Eqs.~(\ref{eq:ConfTT-1}-\ref{eq:ConfTT-2}). Note
408: that we have to set $\tilde M^{ij}$ equal to the trace-free part of the
409: extrinsic curvature.
410:
411: Now suppose we have a guess for a metric and an extrinsic curvature,
412: which ---most likely--- will not satisfy the constraint equations
413: (\ref{eq:HamiltonianConstraintPhysical}) and
414: (\ref{eq:MomentumConstraint}). Set $\tilde\gamma_{ij}$ to the guess
415: for the metric, and set $K$ and $\tilde M^{ij}$ to the trace and
416: trace-free piece of the guess of the extrinsic curvature. By solving
417: the conformal TT equations we can compute $(\gamma_{ij}, K^{ij})$ that
418: satisfy the constraint equations. If our initial guess is ``close''
419: to a true solution, we will have $\psi\approx 1$ and $V^i\approx 0$,
420: so that $\gamma_{ij}$ and $K^{ij}$ will be close to our initial guess.
421:
422: Thus one can guess a metric and extrinsic curvature as well as
423: possible and then solve the conformal TT equations to obtain corrected
424: quantities that satisfy the constraint equations.
425:
426: An artifact of the conformal TT decomposition is that one has no
427: direct handle on the transverse traceless piece with respect to the
428: {\em physical} metric. For any vector $X^i$,
429: \begin{equation}\label{eq:ConformalLong}
430: (\Long X)^{ij}=\psi^{-4}(\tilde\Long X)^{ij}.
431: \end{equation}
432: Thus, Eqs.~(\ref{eq:ConfTT-AijConfAij}) and (\ref{eq:ConfTT-Aij}) imply
433: \begin{equation}\label{eq:ConfTT-Decomposition-psi}
434: A^{ij}=\psi^{-10}\tilde A^{ij}_{TT}+\psi^{-6}(\Long X)^{ij}.
435: \end{equation}
436: For any symmetric traceless tensor $S^{ij}$
437: \begin{equation}
438: \label{eq:DecompositionSymmetricTrace-Free}
439: \nabla_jS^{ij}=\psi^{-10}\tilde\nabla_j\left(\psi^{10}S^{ij}\right).
440: \end{equation}
441: Therefore the first term on the right hand side of
442: Eq.~(\ref{eq:ConfTT-Decomposition-psi}) is transverse-traceless with
443: respect to the physical metric,
444: \begin{equation}
445: \nabla_j\left(\psi^{-10}\tilde A^{ij}_{TT}\right)=0.
446: \end{equation}
447: However, the second term on the right hand side of
448: Eq.~(\ref{eq:ConfTT-Decomposition-psi}) is conformally weighted.
449: Therefore, Eq.~(\ref{eq:ConfTT-Decomposition-psi}) does not represent the
450: usual TT decomposition.
451:
452:
453: %****************************************************************
454: \subsubsection{Physical TT Decomposition}
455: %****************************************************************
456:
457: In this case one decomposes the physical traceless extrinsic
458: curvature directly:
459: \begin{equation}
460: \label{eq:PhysTT-Aij}
461: A^{ij}=A^{ij}_{TT}+(\Long X)^{ij}.
462: \end{equation}
463: As above in the conformal TT decomposition, the momentum constraint
464: becomes an elliptic equation for $X^i$. We can again incorporate the
465: construction of the symmetric transverse traceless tensor
466: $A^{ij}_{TT}$ from a general symmetric tensor $\tilde M^{ij}$ into the
467: momentum constraint. Then one obtains the {\em physical TT equations}:
468:
469: \begin{gather}
470: \label{eq:PhysTT-1}
471: \tilde\nabla^2\psi-\frac{1}{8}\psi\tilde R-\frac{1}{12}\psi^5K^2
472: +\frac{1}{8}\psi^{5}\tilde A_{ij}\tilde A^{ij}=-2\pi G\psi^5\rho,\\
473: \label{eq:PhysTT-2}
474: \tildeLapLong V^i+6(\tilde\Long V)^{ij}\tilde\nabla_j\ln\psi
475: -\frac{2}{3}\tilde\nabla^iK
476: +\psi^{-6}\tilde\nabla_j\tilde M^{ij}=8\pi G\psi^4j^i,
477: \end{gather}
478: where $\tilde A^{ij}$ is defined by
479: \begin{equation}
480: \tilde A^{ij}=(\tilde\Long V)^{ij}+\psi^{-6}\tilde M^{ij}.
481: \end{equation}
482: When we have solved (\ref{eq:PhysTT-1}) and (\ref{eq:PhysTT-2}) for
483: $\psi$ and $V^i$, the physical metric is given by
484: (\ref{eq:ConformalMetric}), and the extrinsic curvature is
485: \begin{equation}
486: \label{eq:PhysTT-3}
487: K^{ij}=\psi^{-4}\left(\tilde A^{ij}+\frac{1}{3}\tilde\gamma^{ij}K\right).
488: \end{equation}
489:
490: We are free to specify the background metric $\tilde\gamma_{ij}$, the
491: trace of the extrinsic curvature $K$, and a symmetric traceless tensor
492: $\tilde M^{ij}$. As with the conformal TT equations, the solution
493: $V^i$ will contain a contribution that removes the longitudinal piece
494: from $\tilde M^{ij}$ and a piece that solves the momentum constraint
495: if $\tilde M^{ij}$ were transverse-traceless.
496:
497: These equations can be used in the same way as the conformal TT
498: equations. Guess a metric and extrinsic curvature, set
499: $\tilde\gamma_{ij}$ to the guess for the metric, and $K$ and $\tilde
500: M^{ij}$ to the trace and trace-free pieces of the guess for the
501: extrinsic curvature. Then solve the physical TT equations to obtain a
502: corrected metric $\gamma_{ij}$ and a corrected extrinsic curvature
503: $K^{ij}$ that satisfy the constraint equations.
504:
505: The transverse traceless piece of $K^{ij}$ (with respect to
506: $\gamma_{ij}$) will be the transverse traceless piece of
507: $\psi^{-10}\tilde M^{ij}$. One can also easily rewrite the physical
508: TT equations such that $\psi^{-10}\tilde M^{ij}$ can be freely chosen
509: instead of $\tilde M^{ij}$. So, in this decomposition we can directly
510: control the TT piece of the physical extrinsic curvature. We have
511: chosen to follow \cite{Cook:2000} since it seems somewhat more natural
512: to specify two conformal quantities, $\tilde\gamma_{ij}$ and $\tilde
513: M^{ij}$ than to specify one conformal and one physical quantity.
514:
515:
516: %****************************************************************
517: \subsubsection{Conformal thin sandwich decomposition}
518: %****************************************************************
519:
520:
521: The conformal and physical TT decompositions rely on a tensor
522: splitting to decompose the trace-free part of the extrinsic curvature.
523: In contrast, the conformal thin sandwich decomposition simply defines
524: $A^{ij}$ by Eq.~(\ref{eq:ConfTT-AijConfAij}) and the decomposition
525: \begin{equation}
526: \label{eq:SandwichTT-3}
527: \tilde A^{ij}\equiv\frac{1}{2\tilde\alpha}
528: \left((\tilde\Long \beta)^{ij}-\tilde u^{ij}\right),
529: \end{equation}
530: where $\tilde u^{ij}$ is symmetric and tracefree.
531: Eq.~(\ref{eq:SandwichTT-3}) is motivated by
532: Eq.~(\ref{eq:Evolution-gamma-TF}): If one evolves an initial-data set
533: with $A^{ij}$ of the form (\ref{eq:SandwichTT-3}) using as lapse and
534: shift
535: \begin{equation}\label{eq:SandwichTT-LapseShift}
536: \begin{aligned}
537: N&=\psi^6\tilde\alpha,\\
538: N^i&=\beta^i,
539: \end{aligned}
540: \end{equation}
541: then
542: \begin{equation}\label{eq:TF-sandwich}
543: % \partial_t\tilde\gamma_{ij}
544: % = \psi^{-4}\TF(\partial_t\gamma_{ij})
545: % =\tilde u_{ij}.
546: \TF(\partial_t\gamma_{ij})=\psi^4\tilde u_{ij}.
547: \end{equation}
548: Therefore, the decomposition (\ref{eq:SandwichTT-3}) is closely
549: related to the kinematical quantities in an evolution. Although
550: $\tilde\alpha$ and $\beta^i$ are introduced in the context of initial
551: data, one usually refers to them as the ``conformal lapse'' and ``shift''.
552: While the form of Eq.~(\ref{eq:SandwichTT-3}) is similar in form to
553: the conformal and physical TT decompositions, there are differences.
554: In particular, $\tilde{u}^{ij}$ is {\em not} divergenceless.
555:
556: Within the {\em conformal thin sandwich decomposition}, the constraint
557: equations take the form:
558: \begin{gather}
559: \label{eq:SandwichTT-1}
560: \tilde\nabla^2\psi-\frac{1}{8}\psi\tilde R-\frac{1}{12}\psi^5K^2
561: +\frac{1}{8}\psi^{-7}\tilde A_{ij}\tilde A^{ij}=-2\pi G\psi^5\rho\\
562: \tildeLapLong\beta^i-(\tilde\Long\beta)^{ij}\tilde\nabla_j\ln\tilde\alpha
563: -\frac{4}{3}\tilde\alpha\psi^6\tilde\nabla^iK
564: \qquad\qquad\qquad\qquad\nonumber\\
565: \label{eq:SandwichTT-2}
566: \qquad\qquad\qquad\qquad -\tilde\alpha\tilde\nabla_j\Big(\frac{1}{\tilde\alpha}\tilde u^{ij}\Big)
567: =16\pi G\tilde\alpha\psi^{10}j^i
568: \end{gather}
569: Having solved Eqs.~(\ref{eq:SandwichTT-1}) and (\ref{eq:SandwichTT-2})
570: for $\psi$ and the vector $\beta^i$, one obtains the physical metric
571: from (\ref{eq:ConformalMetric}) and the extrinsic curvature from
572: \begin{equation}
573: K^{ij}=\psi^{-10}\tilde A^{ij}+\frac{1}{3}\psi^{-4}\tilde\gamma^{ij}K.
574: \end{equation}
575: In this decomposition we are free to specify a conformal metric
576: $\tilde\gamma_{ij}$, the trace of the extrinsic curvature $K$, a
577: symmetric trace-free tensor $\tilde u^{ij}$ and a function $\tilde\alpha$.
578:
579: It seems that the conformal thin sandwich decomposition contains
580: additional degrees of freedom in the form of the function
581: $\tilde\alpha$ and three additional unconstrained components of
582: $\tilde{u}^{ij}$. This is not the case. The longitudinal piece of
583: $\tilde u^{ij}$ corresponds to the gauge choice of the actual shift
584: vector used in an evolution. Thus $\tilde{u}^{ij}$ really only
585: contributes two degrees of freedom, just like $\tilde{M}^{ij}$ in the
586: conformal and physical TT decompositions. Furthermore, we can reach
587: any {\em reasonable} physical solution $(\gamma_{ij}, K^{ij})$ with
588: any {\em reasonable} choice of $\tilde\alpha$; each choice of
589: $\tilde\alpha$ simply defines a new decomposition. A forthcoming
590: article by York\cite{York:2002} will elaborate on these ideas. Note
591: that for $\tilde \alpha=1/2$ we recover the conformal TT
592: decomposition.
593:
594:
595: Let us now turn to the question of how one should pick the freely
596: specifiable data in the conformal thin sandwich approach.
597: We motivate our prescription again by considering how to recover a
598: known spacetime: Assume we are given a full four-dimensional spacetime
599: with 3+1 quantities $\gamma_{0\,ij}$, $K_0^{ij}$, $N_0^i$ and $N_0$.
600: Further assume the spacetime is stationary and the slicing is such that
601: $\partial_t\gamma_{ij}=\partial_tK_{ij}=0$. An example for such a
602: situation is a Kerr black hole in Kerr-Schild or
603: Boyer-Lindquist coordinates.
604:
605: Using $\partial_t\gamma_{0\,ij}=0$ in
606: Eq.~(\ref{eq:Evolution-gamma-TF}) yields a relation for the trace-free
607: extrinsic curvature
608: \begin{equation}
609: A_0^{ij}=\frac{1}{2N_0}(\Long N_0)^{ij}.
610: \end{equation}
611: This is a decomposition of the form (\ref{eq:SandwichTT-3}) with
612: $\tilde u^{ij}=0$. Therefore, if we choose the freely specifiable data
613: for the conformal thin sandwich equations as
614: \begin{equation}
615: \begin{aligned}
616: \tilde\gamma_{ij}&=\gamma_{0\,ij},&
617: \tilde\alpha&=N_0,\\
618: K&=K_0,&\tilde u^{ij}&=0,
619: \end{aligned}
620: \end{equation}
621: and if we use appropriate boundary conditions, then the solution of
622: the conformal thin sandwich equations will be $\psi=1$ and $\beta^i=N_0^i$.
623: As part of the solution, we obtain the shift vector needed for an
624: evolution to produce $\TF(\partial_t\gamma_{ij})=0$.
625: Not needing a guess for the trace-free extrinsic curvature, and having
626: the solution $\beta^i$ automatically provide an initial shift for
627: evolution, make the conformal thin sandwich equations very attractive.
628:
629: In order to generate initial-data slices that permit an evolution with
630: zero time derivative of the conformal metric --- a highly desirable
631: feature for quasi-equilibrium data, or for a situation with holes
632: momentarily at rest --- one can proceed as follows: Set
633: $\tilde\gamma_{ij}$ and $K$ to the guess for the metric and trace of
634: extrinsic curvature, respectively. Set $\tilde\alpha$ to the
635: lapse function that one is going to use in the evolution, and set
636: $\tilde u^{ij}=0$. If these guesses are good, the conformal factor
637: $\psi$ will be close to 1, and $N=\psi^6\tilde\alpha$ as well as
638: $N^i=\beta^i$ give us the actual lapse function and shift vector to use in
639: the evolution.
640:
641: \section{Choices for the freely specifiable data}
642: \label{sec:FreePieces}
643:
644:
645: \subsection{Kerr-Schild coordinates}
646: \label{sec:Choices:Kerr-Schild}
647:
648: We base our choice for the freely specifiable data on a superposition of two
649: Kerr black holes in Kerr-Schild coordinates. In this section we
650: describe the Kerr-Schild solution and collect necessary
651: equations. We also describe how we compute the 3-metric, lapse, shift
652: and extrinsic curvature for a boosted black hole with arbitrary spin.
653:
654: A Kerr-Schild metric is given by
655: \begin{equation}\label{eq:KerrSchild}
656: g_{\mu\nu}=\eta_{\mu\nu}+2Hl_\mu l_\nu,
657: \end{equation}
658: where $\eta_{\mu\nu}$ is the Minkowski metric, and $l_\mu$ is a
659: null-vector with respect to both the full metric and the Minkowski
660: metric: $g^{\mu\nu}l_\mu l_\nu=\eta^{\mu\nu}l_\mu l_\nu=0$. The
661: 3-metric, lapse and shift are
662: \begin{align}
663: \label{eq:KerrSchild-gamma}
664: \gamma_{ij}&=\delta_{ij}+2Hl_il_j,\\
665: N&=(1+2Hl^tl^t)^{-1/2},\\
666: \label{eq:KerrSchild-beta}
667: N^i&=-\frac{2Hl^tl^i}{1+2Hl^tl^t}.
668: \end{align}
669:
670: For a black hole at rest at the origin with mass $M$ and angular
671: momentum $M\vec a$, one has
672: \begin{align}
673: H&=\frac{Mr^3}{r^4+(\vec a\cdot\vec x)^2},\\
674: l^{\mbox{\footnotesize rest}}_\mu&=(1, \vec l_{\mbox{\footnotesize rest}}),\\
675: \vec l_{\mbox{\footnotesize rest}}\;\,&=\frac{r\vec x-\vec a\times\vec x+(\vec a\cdot\vec x)\vec a/r}
676: {r^2+a^2},
677: \end{align}
678: with
679: \begin{equation}
680: r^2=\frac{\vec x^2-\vec a^2}{2}
681: +\left(\frac{(\vec x^2- \vec a^2)^2}{4}
682: +(\vec a\cdot\vec x)^2\right)^{1/2}.
683: \end{equation}
684: For a nonrotating black hole with $\vec a=0$, $H$ has a pole at the
685: origin, whereas for rotating black holes, $r$ has a ring singularity.
686: We will therefore have to excise from the computational domain a
687: region close to the center of the Kerr-Schild black hole.
688:
689: Under a boost, a Kerr-Schild coordinate system transforms into a
690: Kerr-Schild coordinate system. Applying a Lorentz transformation with
691: boost velocity $v^i$ to $l^{\mbox{\footnotesize rest}}_\mu$, we obtain
692: the null-vector $l_\mu$ of the boosted Kerr-Schild coordinate system.
693: Eqs.~(\ref{eq:KerrSchild-gamma}-\ref{eq:KerrSchild-beta}) give then
694: the boosted 3-metric, lapse, and shift. Since all time-dependence is
695: in the uniform motion, evolution with lapse $N$ and shift $N^i$ yields
696: $\partial_t\gamma_{ij}=-v^k\partial_k\gamma_{ij}$, and from
697: Eq.~(\ref{eq:Evolution-gamma}) one can compute the extrinsic curvature
698: \begin{equation}\label{eq:KerrSchild-Kij}
699: K_{ij}=\frac{1}{2N}\left(v^k\partial_k\gamma_{ij}
700: +\nabla_{\!i} N_j+\nabla_{\!j} N_i\right).
701: \end{equation}
702:
703: If this initial-data set is evolved with the shift $N^i$, the black
704: hole will move through the coordinate space with velocity $v^i$.
705: However, if the evolution uses the shift vector $N^i+v^i$, the
706: coordinates will move with the black hole, and the hole will be at
707: rest in coordinate space. The spacetime is nonetheless different from
708: a Kerr black hole at rest. The ADM-momentum will be $P^i_{ADM}=\gamma
709: M v^i$, where $M$ is the rest-mass of the hole and $\gamma=(1-\vec
710: v^2)^{-1/2}$.
711:
712: \subsection{Freely specifiable pieces}
713:
714: We want to generate initial data for a spacetime containing two black
715: holes with masses $M_{\!A,B}$, velocities $\vec v_{\!A,B}$ and spins
716: $M_{\!A}\vec a_A$ and $M_{\!B}\vec a_B$.
717:
718: We follow the proposal of Matzner et al
719: \cite{Matzner-Huq-Shoemaker:1999, Marronetti-Matzner:2000} and base
720: our choices for the freely specifiable choices on two Kerr-Schild
721: coordinate systems describing two individual black holes. The first
722: black hole with label A has an associated Kerr-Schild coordinate
723: system with metric
724: \begin{equation}
725: \label{eq:KerrSchild-HoleA-gamma}
726: \gamma_{A\,ij}=\delta_{ij}+2H_{\!A}\,l_{A\,i}\,l_{A\,j},
727: \end{equation}
728: and with an extrinsic curvature $K_{\!A\,ij}$, a lapse $N_{\!A}$ and a
729: shift $N^i_A$. The trace of the extrinsic curvature is $K_A$. All
730: these quantities can be computed as described in the previous section,
731: \ref{sec:Choices:Kerr-Schild}. The second black hole has a similar set
732: of associated quantities which are labeled with the letter B.
733:
734: For all three decompositions, we need to choose a conformal metric and
735: the trace of the extrinsic curvature. We choose
736: \begin{gather}
737: \label{eq:BinaryKerrSchild-gamma}
738: \tilde\gamma_{ij}=\delta_{ij}+2H_{\!A}\,l_{A\,i}\,l_{A\,j}
739: +2H_{\!B}\,l_{B\,i}\,l_{B\,j}\\
740: K=K_{\!A}+K_B\label{eq:BinaryKerrSchild-K}
741: \end{gather}
742: The metric is singular at the center of each hole. Therefore we have
743: to excise spheres around the center of each hole from the
744: computational domain. We now specify for each decomposition the
745: remaining freely specifiable pieces and boundary conditions.
746:
747: \subsubsection{Conformal TT and physical TT decompositions}
748:
749: For the conformal TT and physical TT decompositions we will be solving
750: for a correction to our guesses.
751: As guess for the trace-free extrinsic curvature, we use a superposition
752: \begin{equation}
753: \label{eq:BinaryKerrSchild-Mij}
754: \tilde M^{ij}=\left(K^{(i}_{A\,k}+K^{(i}_{B\,k}
755: -\frac{1}{3}\delta^{(i}_k(K_A+K_B)\right)\tilde\gamma^{j)k}.
756: \end{equation}
757: $\tilde M^{ij}$ is symmetric and trace-free with respect to the
758: conformal metric, $\tilde\gamma_{ij}\tilde M^{ij}=0$. Solving for a
759: correction only, we expect that $\psi\approx 1$ and $V^i\approx 0$,
760: hence we use Dirichlet boundary conditions
761: \begin{equation}
762: \label{eq:BC-ConfTT-PhysTT}
763: \psi=1, \qquad V^i=0.
764: \end{equation}
765:
766: \subsubsection{Conformal thin sandwich}
767:
768: For conformal thin sandwich, we restrict the discussion to either two
769: black holes at rest, or in a quasi-circular orbit in corotating
770: coordinates. In these cases, one expects small or even vanishing
771: time-derivatives, $\partial_t\gamma_{ij}\approx 0$, and so
772: Eq.~(\ref{eq:TF-sandwich}) yields the simple choice
773: \begin{equation}\label{eq:CTS-uij}
774: \tilde u^{ij}=0.
775: \end{equation}
776:
777: The conformal 3-metric and the trace of the extrinsic curvature are
778: still given by Eqs.~(\ref{eq:BinaryKerrSchild-gamma}) and
779: (\ref{eq:BinaryKerrSchild-K}). Orbiting black holes in a corotating
780: frame will not move in coordinate space, therefore we do not boost the
781: individual Kerr-Schild metrics in this decomposition: $v^i_{A/B}=0$.
782: The lapse functions $N_{A/B}$ and the shifts $N^i_{A/B}$ are also for
783: unboosted Kerr-Schild black holes.
784:
785: We use Dirichlet boundary conditions:
786: \begin{subequations}\label{eq:BC-sandwich}
787: \begin{align}
788: \psi&=1 &&\mbox{all boundaries}\\
789: \label{eq:Sandwich-BC2}
790: \beta^i&=N_{\!A}^i &&\mbox{sphere inside hole A}\\
791: \beta^i&=N_B^i &&\mbox{sphere inside hole B}\\
792: \label{eq:Sandwich-BC4}
793: \beta^i&=\vec\Omega\times\vec r &&\mbox{outer boundary}
794: \end{align}
795: \end{subequations}
796:
797: Eq.~(\ref{eq:Sandwich-BC4}) ensures that we are in a corotating
798: reference frame; the cross-product is performed in flat space, and
799: $\vec\Omega=0$ corresponds to two black holes at rest. Close to the
800: holes we force the shift to be the shift of a single black hole in the
801: hope that this choice will produce a hole that is at rest in
802: coordinate space.
803:
804:
805: For the conformal lapse we use
806: \begin{equation}\label{eq:Sandwich-N1}
807: \tilde\alpha=N_{\!A}+N_B-1
808: \end{equation}
809: or
810: \begin{equation}\label{eq:Sandwich-N2}
811: \tilde \alpha=N_{\!A}\;N_B.
812: \end{equation}
813: The first choice of $\tilde\alpha$ follows the philosophy of adding
814: quantities of each individual hole. However, $\tilde\alpha$ of
815: Eq.~(\ref{eq:Sandwich-N1}) becomes negative sufficiently close to the
816: center of each hole and is therefore a bad choice if the excised
817: spheres are small. The choice (\ref{eq:Sandwich-N2}) does not change
818: sign and has at large distances the same behavior (same $1/r$ term) as
819: (\ref{eq:Sandwich-N1}).
820:
821:
822:
823: \section{Numerical Implementation}
824: \label{sec:Implementation}
825:
826:
827: We implemented an elliptic solver that can solve all three
828: decompositions we described above in complete generality. The solver
829: uses domain decomposition and can handle nontrivial topologies. It is
830: based on pseudospectral collocation, that is, it expresses the
831: solution in each subdomain as an expansion in basis functions. This
832: elliptic solver is described in detail in a separate paper
833: \cite{Pfeiffer-Kidder-etal:2002}.
834:
835: \begin{figure}
836: \includegraphics[scale=0.33, angle=90]{Fig1.eps}
837: \caption{\label{fig:Domains}
838: Structure of domains. Spherical shells around each excised
839: sphere are surrounded by 43 rectangular blocks and another
840: spherical shell. The rectangular blocks touch each other and
841: overlap with all three spherical shells.}
842: \end{figure}
843:
844: From the computational domain we excise two spheres containing the
845: singularities of the Kerr-Schild metric close to the center of each
846: hole. Around each of the excised spheres we place a spherical shell.
847: These shells are patched together with $5\times 3\times 3=45$
848: rectangular blocks, with the two blocks at the location of the spheres
849: removed. Around these 43 blocks, another spherical shell is placed
850: that extends far out, typically to an outer radius of $10^7M$. In the
851: rectangular blocks, we expand in Chebyshev polynomials, while in the
852: spheres we use Chebyshev polynomials radially and spherical harmonics
853: for the angular variables. This setup is depicted in Fig.~\ref{fig:Domains}.
854:
855: The domain decomposition in Fig.~\ref{fig:Domains} is fairly
856: complicated. Even if the shells were made as large as possible, they do
857: not cover the full computational domain when the excised spheres are
858: close together. Thus additional subdomains are needed in any case.
859: Choosing the 43 cubes as depicted allows for relatively small inner
860: shells and for a relatively large inner radius of the outer shell.
861: Thus each shell covers a region of the computational domain in
862: which the angular variations of the solution are fairly low, allowing
863: for comparatively few angular basis-functions.
864:
865: The code can handle a general conformal metric. In principle, the user
866: needs to specify only $\tilde\gamma_{ij}$. Then the code computes
867: $\tilde\gamma^{ij}$, and ---using numerical derivatives--- the Christoffel
868: symbols, Ricci tensor and Riemann scalar. For the special case of the
869: Kerr-Schild metric of a single black hole and the superposition of two
870: Kerr-Schild metrics, Eq.~(\ref{eq:BinaryKerrSchild-gamma}), we compute
871: first derivatives analytically and use numerical derivatives only to
872: compute the Riemann tensor.
873:
874: The solver implements Eqs.~(\ref{eq:ConfTT-1}) and (\ref{eq:ConfTT-2})
875: for the conformal TT decomposition, Eqs.~(\ref{eq:PhysTT-1}) and
876: (\ref{eq:PhysTT-2}) for the physical TT decomposition, and
877: Eqs.~(\ref{eq:SandwichTT-1}) and (\ref{eq:SandwichTT-2}) for the
878: conformal thin sandwich decomposition.
879:
880: After solving for $(\psi, V^i)$ [conformal TT and physical TT], or
881: $(\psi, \beta^i)$ [thin sandwich] we compute the physical metric
882: $\gamma_{ij}$ and the physical extrinsic curvature $K^{ij}$ of the
883: solution. Utilizing these physical quantities $(\gamma_{ij},
884: K^{ij})$, we implement several analysis tools. We evaluate the
885: constraints in the form of
886: Eq.~(\ref{eq:HamiltonianConstraintPhysical}) and
887: (\ref{eq:MomentumConstraint}), we compute ADM-quantities and we search
888: for apparent horizons. Note that these analysis tools are completely
889: independent of the particular decomposition; they rely only on
890: $\gamma_{ij}$ and $K^{ij}$.
891:
892: Next we present tests ensuring that the various systems of equations
893: are solved correctly. We also include tests of the
894: analysis tools showing that we can indeed compute constraints,
895: ADM-quantities and apparent horizons with good accuracy.
896:
897:
898: \subsection{Testing the conformal TT and physical TT decompositions}
899:
900: We can test the solver by conformally distorting a known solution of
901: the constraint equations. Given a solution to the constraint
902: equations $(\gamma_{0\,ij}, K^{ij}_0)$ pick functions
903: \begin{gather}
904: \Psi>0,\qquad W^i
905: \end{gather}
906: and set
907: \begin{align}
908: \label{eq:Test-gamma}
909: \tilde\gamma_{ij}&=\Psi^{-4}\gamma_{0\,ij},\\
910: \label{eq:Test-K}
911: K\;&=K_0,
912: \end{align}
913: and
914: \begin{align}\label{eq:Test-ConfTT-M}
915: \tilde M^{ij}&=\Psi^{10}\left(K_0^{ij}-\frac{1}{3}\gamma_0^{ij}K_0\right)
916: -\Psi^4(\Long_0W)^{ij}
917: \intertext{for conformal TT or}
918: \label{eq:Test-PhysTT-M}
919: \tilde M^{ij}&=\Psi^{10}\left(K_0^{ij}-\frac{1}{3}\gamma_0^{ij}K_0
920: -(\Long_0W)^{ij}\right)
921: \end{align}
922: for physical TT.
923: With these freely specifiable pieces and appropriate boundary
924: conditions, a solution of the conformal TT equations
925: (\ref{eq:ConfTT-1}), (\ref{eq:ConfTT-2}) or the physical TT equations
926: (\ref{eq:PhysTT-1}), (\ref{eq:PhysTT-2}) will be
927: \begin{align}
928: \psi&=\Psi\\
929: V^i&=W^i.
930: \end{align}
931: From Eq.~(\ref{eq:ConformalMetric}) we recover our initial metric
932: $\gamma_{0\,ij}$, and from Eq.~(\ref{eq:ConfTT-3}) [conformal TT] or
933: Eq.~(\ref{eq:PhysTT-3}) [physical TT] we recover the extrinsic
934: curvature $K^{ij}_0$.
935:
936: In our tests we used the particular choices
937: \begin{gather}
938: \label{eq:Psi}
939: \Psi=1+\frac{8(r-2)}{36+x^2+0.9y^2+1.3(z-1)^2}\\
940: \label{eq:W^i}
941: W^i=\frac{50(r-2)}{(6^4+r^4)} (-y, x, 1).
942: \end{gather}
943:
944: \begin{figure}
945: \centerline{\includegraphics[scale=0.38]{Fig2.eps}}
946: \caption{\label{fig:Testing-PlotDistortion}
947: Plot of the functions $\Psi$ and $W^i$ from Eqs.~(\ref{eq:Psi})
948: and (\ref{eq:W^i}). The solid line depicts $\Psi$ along the
949: positive $x$-axis, the short dashed line depicts $\Psi$ along the
950: negative $z$-axis. The long dashed line is a plot of $W^y$ along the
951: positive $x$-axis.}
952: \end{figure}
953:
954: These functions are plotted in Fig.~\ref{fig:Testing-PlotDistortion}.
955: $\Psi$ varies between 0.8 and 1.5, $W^i$ varies between $\pm 0.5$, and
956: both take their maximum values around distance $\sim 7$ from the
957: center of the hole. We used for $(\tilde\gamma_{0\,ij}, K^{ij}_0)$ a
958: single, boosted, spinning black hole in Kerr-Schild coordinates.
959:
960: \begin{figure}
961: % this pic produced with /home/harald/Eric/Run/Neville/01Jul/05E/PlotConv.sm
962: \centerline{\includegraphics[scale=0.38]{Fig3.eps}}
963: \caption{\label{fig:Testing-ConfTT}
964: Testing the solver for the conformal TT decomposition.
965: Eqs.~(\ref{eq:ConfTT-1}-\ref{eq:ConfTT-2}) with freely specifiable
966: data given by Eqs.~(\ref{eq:Test-gamma}-\ref{eq:Test-ConfTT-M}) are
967: solved in a single spherical shell with $1.5M<r<10M$. $N$ is the
968: cube root of the total number of unknowns. Plotted are the L2-norms
969: of $\psi-\Psi$, $V^x-W^x$, and the residuals of Hamiltonian and
970: momentum constraints, Eqs.~(\ref{eq:HamiltonianConstraintPhysical})
971: and (\ref{eq:MomentumConstraint}). }
972: \end{figure}
973:
974: Figure~\ref{fig:Testing-ConfTT} shows results of testing the conformal
975: TT decomposition on a single spherical shell. The numerical solution
976: $(\psi, V^i)$ converges to the analytic solutions $(\Psi, W^i)$
977: exponentially with the number of basis functions as expected for a
978: properly constructed spectral method. Moreover, the reconstructed
979: metric and extrinsic curvature satisfy the constraints.
980:
981: \begin{figure}
982: % plot produced by /home/harald/Eric/Run/SP2/01Jul/07E/PlotConv.sm
983: \centerline{\includegraphics[scale=0.38]{Fig4.eps}}
984: \caption{\label{fig:Testing-PhysTT}
985: Physical TT decomposition with domain decomposition.
986: Eqs.~(\ref{eq:PhysTT-1}), (\ref{eq:PhysTT-2}) with freely
987: specifiable data given by Eq.~(\ref{eq:Test-gamma}, \ref{eq:Test-K},
988: \ref{eq:Test-PhysTT-M}) are solved in multiple domains (one inner
989: spherical shell, 26 rectangular blocks, one outer spherical shell).
990: $N$ is the cube root of the total number of grid-points. {\em diff}
991: denotes the L2-norm of the difference of the solution and the
992: solution at next lower resolution. Triangles denote the L2-norm of
993: the difference to the analytic solution. The remaining symbols
994: denote the errors of numerically extracted ADM-quantities. }
995: \end{figure}
996:
997: Now we test the solver for the physical TT decomposition, and
998: demonstrate that we can correctly deal with multiple domains. In this
999: example the computational domain is covered by an inner spherical
1000: shell extending for $1.5M\le r\le 10M$. This shell is surrounded by
1001: 26 rectangular blocks that overlap with the shell and extend out to
1002: $x,y,z=\pm 25M$. Finally another spherical shell covers the region
1003: $20M\le r\le 10^6M$. As can be seen in Fig.~\ref{fig:Testing-PhysTT},
1004: the solution converges again exponentially.
1005:
1006: For realistic cases we do not know the analytic solution and therefore
1007: need a measure of the error. Our major tool will be the change in
1008: results between different resolution. In particular we consider the
1009: $L_2$ norm of the point-wise differences of the solution at some resolution
1010: and at the next lower resolution. This diagnostic is labeled by circles
1011: in Fig.~\ref{fig:Testing-PhysTT}. Since the solution converges
1012: exponentially, these circles essentially give the error of the {\em
1013: lower} of the two resolutions.
1014:
1015:
1016: In addition to testing the equations, this example tests domain
1017: decomposition and the integration routines for the ADM quantities.
1018: The ADM quantities are computed by the standard integrals at infinity
1019: in Cartesian coordinates,
1020: \begin{align}\label{eq:EADM}
1021: E_{ADM}&=\frac{1}{16\pi}\int_{\infty}
1022: \left(\gamma_{ij,j}-\gamma_{jj,i}\right)\,d^2S_i,\\
1023: \label{eq:JADM}
1024: J_{(\xi)}&=\frac{1}{8\pi}\int_{\infty}
1025: \left(K^{ij}-\gamma^{ij}K\right)\xi_j\,d^2S_i.
1026: \end{align}
1027: For the $x$-component of the linear ADM-momentum, $\xi=\hat e_x$ in
1028: Eq.~(\ref{eq:JADM}). The choice $\xi=x\hat e_y-y\hat e_x$ yields the
1029: $z$-component of the ADM-like angular momentum as defined by York
1030: \cite{York:1979}. Since the space is asymptotically flat
1031: there is no distinction between upper and lower indices in
1032: Eqs.~(\ref{eq:EADM}) and (\ref{eq:JADM}). Note that
1033: Eq.~(\ref{eq:EADM}) reduces to the familiar monopole term
1034: \begin{equation}
1035: -\frac{1}{2\pi}\int_{\infty}\partial_r\psi\,dA
1036: \end{equation}
1037: {\em only} for quasi-isotropic coordinates.
1038: Our outer domain is large, but since
1039: it does not extend to infinity, we extrapolate $r\to\infty$.
1040:
1041: For a Kerr black hole with mass $M$ and spin $\vec a$, that is boosted
1042: to velocity $\vec v$, the ADM-quantities will be
1043: \begin{align}
1044: E_{ADM}&=\gamma M,\\
1045: \vec P_{ADM}&=\gamma M\vec v,\\
1046: \label{eq:J-boostedBH}
1047: \vec J_{ADM}&=\left[\gamma\vec a-(\gamma-1)\frac{(\vec a\,\vec v)\,\vec v}{\vec v\,^2}\right]M,
1048: \end{align}
1049: where $\gamma=(1-\vec v\,^2)^{-1/2}$ denotes the Lorentz factor.
1050: Eq.~(\ref{eq:J-boostedBH}) reflects the fact that under a boost, the
1051: component of the angular momentum perpendicular to the boost-direction
1052: is multiplied by $\gamma$.
1053:
1054: The example in Fig.~\ref{fig:Testing-PhysTT} uses
1055: %$M=1.2$,
1056: $\vec v=(0.2, 0.3, 0.4)$, and $\vec a=(-1/4, 1/4, 1/6)M$. The
1057: evaluation of the angular momentum $J_z$ seems to be less accurate
1058: since our current procedure to extrapolate to infinity magnifies
1059: roundoff. We plan to improve this in a future version of the code.
1060: Until then we seem to be limited to an accuracy of $\sim 10^{-6}$.
1061:
1062:
1063: %****************************************************************
1064: \subsection{Testing conformal thin sandwich equations}
1065: %****************************************************************
1066:
1067: The previous decompositions could be tested with a conformally
1068: distorted known solution. In order to test the conformal thin
1069: sandwich decomposition we need to find an analytic decomposition of
1070: the form (\ref{eq:SandwichTT-3}). To do this, we
1071: start with a stationary solution of Einstein's equations and boost it with
1072: uniform velocity $v^i$. Denote the metric, extrinsic curvature, lapse
1073: and shift of this boosted spacetime by $\tilde\gamma_{0\,ij}$,
1074: $K^{ij}_0=A^{ij}_0 +\frac{1}{3}\gamma^{ij}_0K_0$, $N_0$ and $N^i_0$,
1075: respectively. Since we boosted the static solution, we will {\em not}
1076: find $\partial_t\gamma_{ij}=0$ if we evolve it with the shift $N^i_0$.
1077: However, all time-dependence of this spacetime is due to the uniform
1078: motion, so in the comoving reference frame specified by the
1079: shift $N^i_0+v^i$, we will find $\partial_t\gamma_{ij}=0$. In this
1080: case, Eq.~(\ref{eq:Evolution-gamma-TF}) yields
1081: \begin{equation}\label{eq:Aij-boost}
1082: A^{ij}_0=\frac{1}{2N_0}(\Long(N_0+v))^{ij}.
1083: \end{equation}
1084: If we choose $\tilde \alpha=N_0$ and
1085: $\tilde u^{ij}=0$, the thin sandwich equations (\ref{eq:SandwichTT-1})
1086: and (\ref{eq:SandwichTT-2}) will thus be solved by $\psi=1$ and
1087: $\beta^i=N^i_0+v^i$. Similar to the conformal TT and physical TT
1088: decomposition above, we can also conformally distort the metric
1089: $\gamma_{0\,ij}$. Furthermore, we can consider nonvanishing $\tilde
1090: u^{ij}$. We arrive at the following method to test the solver for the
1091: conformal thin sandwich equations:
1092:
1093: Given a boosted version of a stationary solution with shift $N_0^i$,
1094: lapse $N_0$, 3-metric $\gamma_{0\,ij}$, trace of extrinsic curvature
1095: $K_0$, and boost-velocity $v^i$. Pick any functions
1096: \begin{gather}
1097: \Psi>0\\
1098: W^i
1099: \end{gather}
1100: and set
1101: \begin{align}
1102: \label{eq:Test-Sandwich-gamma}
1103: \tilde\gamma_{ij}&=\Psi^{-4}\gamma_{0\,ij}\\
1104: K\;&=K_0\\
1105: \tilde\alpha\;\,&=\Psi^{-6}N_0\\
1106: \label{eq:Test-Sandwich-u}
1107: \tilde u^{ij}&=\Psi^4(\Long_0 W)^{ij}
1108: \end{align}
1109: Then a solution to the thin sandwich equations
1110: (\ref{eq:SandwichTT-1}-\ref{eq:SandwichTT-2}) will be
1111: \begin{align}
1112: \psi&=\Psi\label{eq:psi-static-boosted}\\
1113: \beta^i&=N_0^i+v^i+W^i\label{eq:beta-static-boosted}
1114: \end{align}
1115: assuming boundary conditions respecting this solution.
1116:
1117: \begin{figure}
1118: % plot generated by /home/harald/Run/Neville/01Jul/14A/PlotConv.sm
1119: \centerline{\includegraphics[scale=0.38]{Fig5.eps}}
1120: \caption{\label{fig:Testing-Sandwich}
1121: Testing thin sandwich decomposition with apparent horizon searches.
1122: Equations~(\ref{eq:SandwichTT-1}) and (\ref{eq:SandwichTT-2}) with
1123: freely specifiable data given by
1124: Eqs.~(\ref{eq:Test-Sandwich-gamma})--(\ref{eq:Test-Sandwich-u}) are solved
1125: in a single spherical shell. $N$ and {\em diff} as in
1126: Fig.~\ref{fig:Testing-PhysTT}. Apparent horizon searches with different
1127: surface expansion order $L$ were performed, and the errors of the
1128: apparent horizon mass $M_{AH}$ are plotted. }
1129: \end{figure}
1130:
1131: Figure~\ref{fig:Testing-Sandwich} shows results of this test for a
1132: single spherical shell and a Kerr black hole with $\vec v=(0.2,
1133: -0.3, 0.1)$, $\vec a=(0.4, 0.3, 0.1)M$. The solution converges to the
1134: expected analytical result exponentially. In addition, apparent
1135: horizon searches were performed. For the numerically found apparent
1136: horizons, the apparent horizon area $A_{AH}$ as well as the apparent
1137: horizon mass
1138: \begin{equation}
1139: \label{eq:M_AH}
1140: M_{AH}=\sqrt{\frac{A_{AH}}{16\pi}}
1141: \end{equation}
1142: were computed. The figure compares $M_{AH}$ to the expected value
1143: \begin{equation}
1144: M\left(\frac{1}{2}+\frac{1}{2}\sqrt{1-\frac{\vec a^2}{M^2}}\;\right)^{1/2}.
1145: \end{equation}
1146: As described in \cite{Baumgarte-Cook-etal:1996,
1147: Pfeiffer-Teukolsky-Cook:2000}, the apparent horizon finder expands
1148: the apparent horizon surface in spherical harmonics up to a fixed
1149: order $L$. For fixed $L$, the error in the apparent horizon mass is
1150: dominated by discretization error of the elliptic solver at low
1151: resolution $N$. As $N$ is increased, the discretization error of the
1152: elliptic solver falls below the error due to finite $L$. Then the
1153: error in $M_{AH}$ becomes independent of $N$. Since the expansion in
1154: spherical harmonics is {\em spectral}, the achievable resolution
1155: increases exponentially with $L$. Note that for exponential
1156: convergence it is necessary to position the rays in the apparent
1157: horizon finder at the abscissas of Gauss-Legendre integration.
1158:
1159:
1160: \subsection{Convergence of binary black hole solutions}
1161:
1162: \begin{figure}
1163: % /home1/harald/Eric/Run/SP2/01Jul/09F/PlotConv.sm
1164: \centerline{\includegraphics[scale=0.38]{Fig6.eps}}
1165: \caption{\label{fig:Convergence-09F}
1166: Binary black hole with conformal TT decomposition. The residuals
1167: of several quantities are plotted as a function of the cube root
1168: of the total number of grid points. {\em diff} as in
1169: Fig.~\ref{fig:Testing-PhysTT}, {\em Ham} and {\em Mom} are the
1170: residuals of Hamiltonian and momentum constraints. $E_{ADM}$
1171: denotes the difference between ADM-energy at resolution $N$ and
1172: ADM-energy at highest resolution.}
1173: \end{figure}
1174:
1175: Figure~\ref{fig:Convergence-09F} present the convergence of the solver
1176: in the binary black hole case. In this particular example, the
1177: conformal TT equations were solved for two black holes at rest with
1178: coordinate separation of $10M$. The computational domain is structured
1179: as in Fig.~\ref{fig:Domains}. The excised spheres have radius
1180: $r_{exc}=2M$, the inner spherical shells extend to radius $4M$. The
1181: rectangular blocks cover space up to $x,y,z=\pm 25M$, and the outer
1182: spherical shell extending from inner radius $20M$ to an outer radius
1183: of $R=10^7M$.
1184:
1185: We do not use fall-off boundary conditions at the outer boundary; we
1186: simply set $\psi=1$ there. This limits the computations presented in
1187: this paper to an accuracy of order $1/R\sim 10^{-7}$.
1188: Figure~\ref{fig:Convergence-09F} shows that even for the next to
1189: highest resolution ($N\approx 80$) the solution will be limited by the
1190: outer boundary condition. All results presented in the following
1191: section are obtained at resolutions around $N\approx 80$. If the need
1192: arises to obtain solutions with higher accuracy, one can easily change
1193: to a fall-off boundary condition, or just move the outer boundary
1194: further out.
1195:
1196:
1197: %****************************************************************
1198: \section{Results}
1199: \label{sec:Results}
1200: %****************************************************************
1201:
1202: The purpose of this paper is to compare the initial-data sets
1203: generated by different decompositions using simple choices for
1204: the freely specifiable pieces in each decomposition. We solve
1205:
1206: $\bullet$ {\bf ConfTT:} Conformal TT equations (\ref{eq:ConfTT-1})
1207: and (\ref{eq:ConfTT-2}) with freely specifiable pieces and boundary
1208: conditions given by Eqs.~(\ref{eq:BinaryKerrSchild-gamma}),
1209: (\ref{eq:BinaryKerrSchild-K}), (\ref{eq:BinaryKerrSchild-Mij})
1210: and (\ref{eq:BC-ConfTT-PhysTT}).
1211:
1212: $\bullet$ {\bf PhysTT:} Physical TT equations (\ref{eq:PhysTT-1})
1213: and (\ref{eq:PhysTT-2}) with same freely specifiable pieces and
1214: boundary conditions as ConfTT.
1215:
1216: $\bullet$ {\bf CTS:} Conformal thin sandwich
1217: equations (\ref{eq:SandwichTT-1}) and (\ref{eq:SandwichTT-2}) with
1218: freely specifiable pieces and boundary conditions given by
1219: Eqs.~(\ref{eq:BinaryKerrSchild-gamma}), (\ref{eq:BinaryKerrSchild-K}),
1220: (\ref{eq:CTS-uij}) and (\ref{eq:BC-sandwich}). The lapse
1221: $\tilde\alpha$ is given by either Eq.~(\ref{eq:Sandwich-N1}), or by
1222: Eq.~(\ref{eq:Sandwich-N2}).
1223:
1224: We will apply the terms ``ConfTT'', ``PhysTT'' and ``CTS'' only to
1225: these particular choices of decomposition, freely specifiable pieces
1226: and boundary conditions. When referring to different freely
1227: specifiable pieces, or a decomposition in general, we will not use
1228: these shortcuts. If we need to distinguish between the two choices of
1229: $\tilde\alpha$ in CTS, we will use ``CTS-add'' for the additive lapse
1230: Eq.~(\ref{eq:Sandwich-N1}) and ``CTS-mult''for the multiplicative
1231: lapse Eq.~(\ref{eq:Sandwich-N2}).
1232: Below in section \ref{sec:mConfTT} we will also introduce as a forth
1233: term ``mConfTT''.
1234:
1235: \subsection{Binary black hole at rest}
1236: \label{sec:ResultsBHatRest}
1237:
1238:
1239: We first examine the simplest possible configuration: Two black holes
1240: at rest with equal mass, zero spin, and with some fixed proper
1241: separation between the apparent horizons of the holes. We solve
1242: \begin{itemize}
1243: \item ConfTT
1244: \item PhysTT
1245: \item CTS (with both choices of $\tilde\alpha$).
1246: \end{itemize}
1247: In the comparisons, we also include inversion symmetric conformally
1248: flat initial data obtained with the conformal-imaging formalism.
1249:
1250: We excised spheres with radius $r_{exc}=2M$, which is close to the
1251: event horizon for an individual Eddington-Finkelstein black hole.
1252: This results in the boundary conditions being imposed close to, but
1253: within the apparent horizons of the black holes.
1254: The centers of the excised spheres have
1255: a coordinate separation of $d=10M$.
1256:
1257: We now discuss the solutions. The conformal factor $\psi$ is very
1258: close to $1$ for each of the three decompositions. It deviates from
1259: $1$ by less than 0.02, indicating that a conformal metric based of a
1260: superposition of Kerr-Schild metrics does not deviate far from the
1261: constraint surface.
1262:
1263: \begin{figure}
1264: % created by Neville:~/DataEric/Run/SP2/PlotCuts.sm
1265: \centerline{\includegraphics[scale=.38]{Fig7.eps}}
1266: \caption{\label{fig:Compare-cuts}
1267: The conformal factor $\psi$ along the axis connecting the holes for
1268: several decompositions. $x$ measures the distance from the center of
1269: mass, so that the excised sphere is located between $3<x<7$.
1270: mConfTT is explained below in section \ref{sec:mConfTT}. The
1271: solution of PhysTT is not plotted since it is within the line
1272: thickness of ConfTT. The insert shows an enlargement for large $x$.
1273: }
1274: \end{figure}
1275:
1276: Figure~\ref{fig:Compare-cuts} presents a plot of the conformal factor
1277: along the axis through the centers of the holes. One sees that $\psi$
1278: is close to $1$; however, between the holes ConfTT and CTS force
1279: $\psi$ in {\em opposite} directions. For CTS, $\psi>1$ between the
1280: holes, for ConfTT, $\psi<1$! The contour plots in Fig.~\ref{fig:Contours}
1281: also show this striking difference between the
1282: decompositions.
1283:
1284: \begin{figure}
1285: \centerline{
1286: \includegraphics[scale=0.23]{Fig8a.eps}
1287: \includegraphics[scale=0.23]{Fig8b.eps}
1288: }
1289: \caption{\label{fig:Contours}
1290: Black holes at rest: Contour plots of the conformal factor $\psi$
1291: for ConfTT (left) and CTS-add (right). The circles denote the
1292: excised spheres of radius $2$. }
1293: \end{figure}
1294:
1295: \begin{table*}
1296: \caption{\label{tab:EADM}
1297: Solutions of different decompositions for two black holes at rest.
1298: Ham and Mom are the rms residuals of the Hamiltonian
1299: and momentum constraint, $\ell$ is the proper separation between
1300: the apparent horizons. {\em mConfTT} represents the modified conformal TT
1301: decomposition which is explained in section \ref{sec:mConfTT}.
1302: {\em inv. symm.} represents a conformally flat, time symmetric
1303: and inversion symmetric solution of the Hamiltonian constraint.
1304: }
1305: \begin{ruledtabular}
1306: \begin{tabular}{ccccccccccc}
1307: & Ham & Mom & $E_{ADM}$ & $A_{AH}$ & $M_{AH}$ & $\ell$ & $\ell/m$ & $E_{ADM}/m$ & $E_{MPRC}/E_{ADM}$ & $E_b/\mu$
1308: \rule[-.65em]{0.em}{1.3em}\\\hline
1309: % ConfTT is 01Jul/09F
1310: ConfTT & $9\times 10^{-7}$ & $4\times 10^{-7}$ & 2.06486 & 57.7369 & 1.07175 & 8.062 & 3.761 & 0.9633 & 0.2660 & -0.1467\\
1311: % PhysTT is 01Jul/16A
1312: PhysTT & $9\times 10^{-7}$ & $3\times 10^{-7}$ & 2.06490 & 57.7389 & 1.07176 & 8.062 & 3.761 & 0.9633 & 0.2660 & -0.1467\\
1313: % Sandwich is 01Jun/27B, postprocessed on neville
1314: CTS-add & $2\times 10^{-6}$ & $4\times 10^{-7}$ & 2.08121 & 62.3116 & 1.11340 & 8.039 & 3.610 & 0.9346 & 0.2434 & -0.2615\\ \hline
1315: % Sandwich mult is 01Nov/10A
1316: CTS-mult & $2\times 10^{-6}$ & $5\times 10^{-7}$ & 2.05851 & 60.8113 & 1.09991 & 8.080 & 3.672 & 0.9358 & 0.2444 & -0.2569\\
1317: % mConfTT is 01Nov/04B
1318: mConfTT & $3\times 10^{-6}$ & $1\times 10^{-6}$ & 2.0827 & 62.404 & 1.1142 &&& 0.9346 & 0.2434 & -0.2617\\ \hline
1319: inv. symm. & - & - & 4.36387 & 284.851 & 2.38053 & 17.731 & 3.724 & 0.9166 & 0.2285 & -0.3337
1320: \end{tabular}
1321: \end{ruledtabular}
1322: \end{table*}
1323:
1324:
1325: The result of PhysTT was found to be almost identical with ConfTT.
1326: This is reasonable, since these two decompositions differ only in that
1327: in one case the TT decomposition is with respect to the conformal
1328: metric, and in the other case the TT decomposition it is with respect
1329: to the physical metric. Since $\psi\approx 1$, the conformal metric is
1330: almost identical to the physical metric, and only minor differences
1331: arise. In the following we will often use ConfTT/PhysTT when referring
1332: to both data sets.
1333:
1334: We performed apparent horizon searches for these cases. For all
1335: data sets, the apparent horizon is outside the sphere with radius
1336: $2M$, that is outside the coordinate location for the apparent horizon
1337: in a single hole spacetime.
1338: For ConfTT/PhysTT, the radius of the apparent horizon surface is
1339: $\approx 2.05M$, for CTS it is $\approx 2.15M$. We computed
1340: the apparent horizon area $A_{AH}$, the apparent horizon mass
1341: \begin{equation}
1342: \label{eq:MAH}
1343: M_{AH}=\sqrt{\frac{A_{AH}}{16\pi}}
1344: \end{equation}
1345: of either hole, and the combined mass of both holes,
1346: \begin{equation}\label{eq:m}
1347: m=2M_{AH}.
1348: \end{equation}
1349: There is no rigorous definition of the mass of an individual black
1350: hole in a binary black hole spacetime, and Eq.~(\ref{eq:MAH})
1351: represents the true mass on an individual black hole
1352: only in the limit of wide separation of the black holes.
1353: A hard upper limit on the possible gravitational radiation
1354: emitted to infinity during the coalescence process of a binary
1355: will be
1356: \begin{equation}\label{eq:MPRC}
1357: E_{MPRC}=E_{ADM}-\sqrt{\frac{2A_{AH}}{16\pi}},
1358: \end{equation}
1359: where $2A_{AH}$ is the combined apparent horizon area of both holes.
1360: Thus, $E_{MPRC}$ represents the maximum possible radiation content
1361: (MPRC) of the initial data. This, of course, makes the unlikely
1362: assumption that the binary radiates away all of its angular momentum.
1363:
1364: We also compute the proper separation~$\ell$ between the apparent
1365: horizon surfaces along the straight line connecting the centers of the excised
1366: spheres. In order to compare different data sets we consider the
1367: dimensionless quantities $\ell/E_{ADM}$, $E_{ADM}/m$ and
1368: $E_{MPRC}/E_{ADM}$. We will also use $E_b/\mu$ which will be defined
1369: shortly.
1370:
1371: Table~\ref{tab:EADM} summarizes these quantities for all three
1372: decompositions. It also includes results for inversion symmetric
1373: initial data, which for black holes at rest reduces to the Misner
1374: data\cite{Misner:1963}\footnote{Although the Misner solution can be
1375: obtained analytically, we found it more convenient to solve the
1376: Hamiltonian constraint numerically. The configuration in
1377: Table~\ref{tab:EADM} corresponds to a separation $\beta=12$ in terms of
1378: \cite{Cook:1991}.}. The results in Table~\ref{tab:EADM} are
1379: intended to represent nearly the {\em same} physical configuration.
1380:
1381: From Table~\ref{tab:EADM}, one finds that the black holes have roughly
1382: the same dimensionless proper separation. However, the scaled
1383: ADM-energy $E_{ADM}/m$ differs by as much as 4.7\% between the
1384: different data sets. $E_{MPRC}/E_{ADM}$, which does not depend on any
1385: notion of individual black hole masses at all, differs by 16\% between
1386: the different data sets.
1387:
1388: The inversion symmetric data has lowest $E_{ADM}/m$ and
1389: $E_{MPRC}/E_{ADM}$, CTS has somewhat larger values, and ConfTT/PhysTT
1390: lead to the biggest values for $E_{ADM}/m$ and $E_{MPRC}/E_{ADM}$.
1391: This indicates that, relative to the sizes of the black holes,
1392: ConfTT/PhysTT and CTS probably contain some excess energy.
1393:
1394: A slightly different argument uses the binding energy which is defined
1395: as
1396: \begin{equation}
1397: \frac{E_b}{\mu}\equiv\frac{E_{ADM}-2M_{AH}}{\mu},
1398: \end{equation}
1399: where $\mu=M_{AH}/2$ is the reduced mass. Two Newtonian point masses
1400: at rest satisfy
1401: \begin{equation}\label{eq:Eb-Newton}
1402: \frac{E_b}\mu = -\frac{1}{\ell/m}.
1403: \end{equation}
1404: From Table~\ref{tab:EADM} we see that for ConfTT/PhysTT, $|E_b/\mu| >
1405: (l/m)^{-1}$, and for CTS, $|E_b/\mu|\approx (l/m)^{-1}$. Since gravity
1406: in general relativity is typically {\em stronger} than Newtonian
1407: gravity, we find again that CTS and ConfTT/PhysTT contain too much
1408: energy relative to the black hole masses, ConfTT/PhysTT having even
1409: more than CTS.
1410:
1411: The proper separation between the apparent horizons $\ell/m$ is about
1412: 4\% smaller for CTS than for ConfTT/PhysTT. By
1413: Eq.~(\ref{eq:Eb-Newton}) this should lead to a relative difference in binding
1414: energy of the same order of magnitude. Since $E_b/\mu$ differs by
1415: almost a factor of two between the different decompositions, the
1416: differences in $\ell/m$ play only a minor role.
1417:
1418: %****************************************************************
1419: \subsection{Configurations with angular momentum}
1420: \label{sec:resultOrbiting}
1421: %****************************************************************
1422:
1423: Now we consider configurations which are approximating two black holes
1424: in orbit around each other. The conformal metric is still a
1425: superposition of two Kerr Schild metrics. The black holes are located
1426: along the $x$-axis with a coordinate separation $d$. For
1427: ConfTT/PhysTT, we boost the individual holes to some velocity $\pm
1428: v\hat e_y$ along the $y$-axis. For CTS we go to a co-rotating frame
1429: with an angular frequency $\vec{\Omega}=\Omega\hat e_z$. Thus, for
1430: each decomposition we have a two parameter family of solutions, the
1431: parameters being $(d,v)$ for ConfTT and PhysTT, and
1432: $(d,\Omega)$ for CTS.
1433:
1434: By symmetry, these configuration will have an ADM angular momentum
1435: parallel to the $z$-axis which we denote by $J$. In order to compare
1436: solutions among each other, and against the conformally flat inversion
1437: symmetric data sets, we adjust the parameters $(d,v)$ and
1438: $(d,\Omega)$, such that each initial-data set has angular momentum
1439: $J/\mu m=2.976$ and a proper separation between the apparent horizons
1440: of $l/m=4.880$. In Ref.~\cite{Cook:1994}, these values were determined to be
1441: the angular momentum and proper separation of a binary black hole at
1442: the innermost stable circular orbit.
1443:
1444: \begin{table*}
1445: \caption{\label{tab:EADM2}
1446: Initial-data sets generated by different decompositions for binary
1447: black holes with the same angular momentum $J/\mu m$ and separation $\ell/m$.
1448: The mConfTT dataset is explained in section \ref{sec:mConfTT}. It should
1449: be compared to CTS-add.
1450: }
1451: \begin{ruledtabular}
1452: \begin{tabular}{ccccccccc}
1453: & parameters & $M_{AH}$ & $E_{ADM}$ & $J/\mu m$ & $\ell/m$ & $E_{ADM}/m$ & $E_{MPRC}/E_{ADM}$ & $E_b/\mu$
1454: \rule[-.65em]{0.em}{1.3em}\\\hline
1455: % ConfTT is /Eric/Run/SP2/01Okt/20B
1456: ConfTT & $d=11.899, v=0.26865$ & 1.06368 & 2.12035 & 2.9759 & 4.879 & 0.9967 & 0.2906 & -0.0132\\
1457: % PhysTT is /Eric/Run/SP2/01Nov/05E
1458: PhysTT & $d=11.899, v=0.26865$ & 1.06369 & 2.12037 & 2.9757 & 4.879 & 0.9967 & 0.2906 & -0.0132\\
1459: % sandwich4 is Eric/Run/SP2/01Okt/23A
1460: CTS-add & $d=11.860, \Omega=0.0415$ & 1.07542 & 2.10391 & 2.9789 & 4.884 & 0.9782 & 0.2771 &-0.0873\\
1461: % sandwich4 is Eric/Run/SP2/01Nov/09D
1462: CTS-mult & $d=11.750, \Omega=0.0421$ & 1.06528 & 2.08436 & 2.9776 & 4.880 & 0.9783 & 0.2772 & -0.0867\\
1463: \hline
1464: mConfTT & $d=11.860, \Omega=0.0415$ & 1.0758 & 2.1061 & 3.011 & 4.883 & 0.979 & 0.278 & -0.085\\
1465: %mConfTT-inert & $d=11.899, v=0.26865$ & 1.1310 & 2.2006 & 2.592 & 4.578 & 0.973 & 0.273 & -0.108 \\\hline
1466: inv. symm.\footnote{Data taken from \cite{Cook:1994}} & & & & 2.976 & 4.880 & 0.9774 & 0.2766 & -0.09030
1467: \end{tabular}
1468: \end{ruledtabular}
1469: \end{table*}
1470:
1471: Table~\ref{tab:EADM2} lists the parameters corresponding to this
1472: situation as well as results for each initial-data set\footnote{
1473: Because of the Lorentz contraction, the apparent horizons for
1474: ConfTT/PhysTT intersect the sphere with radius $2$. In order to have
1475: the full apparent horizon inside the computational domain, the
1476: radius of the excised spheres was reduced to 1.9 for these data
1477: sets.}. As with the configuration with black holes at rest, we find
1478: again that ConfTT/PhysTT and CTS lead to different ADM-energies.
1479: Now, $E_{ADM}/m$ and $E_{MPRC}/E_{ADM}$ differ by $0.02$ and $0.013$,
1480: respectively, between CTS and ConfTT/PhysTT. However, in contrast to
1481: the cases where the black holes at rest, now CTS and the inversion symmetric data set
1482: have very similar values for $E_{ADM}/m$ and $E_{MPRC}/E_{ADM}$.
1483:
1484: %****************************************************************
1485: \subsection{Reconciling conformal TT and thin sandwich}
1486: %****************************************************************
1487: \label{sec:mConfTT}
1488:
1489: We now investigate further the difference between ConfTT/PhysTT and
1490: CTS. Since the resulting initial-data sets for PhysTT and ConfTT are
1491: very similar, we restrict our discussion to ConfTT.
1492:
1493: \subsubsection{Motivation}
1494:
1495: The construction of binary black hole data for the ConfTT/PhysTT
1496: cases produces an extrinsic curvature that almost certainly
1497: contains a significant TT component. It would be interesting to know
1498: how significant this component is to the value of the various physical
1499: parameters we are comparing. Ideally, we would like to completely
1500: eliminate the TT component and see what effect this has on the resulting
1501: data sets. Unfortunately, this is a difficult, if not impossible,
1502: task.
1503:
1504: The TT component of a symmetric tensor $M^{ij}$ is defined as
1505: \begin{equation}
1506: M^{ij}_{TT} \equiv M^{ij} - (\Long Y)^{ij},
1507: \end{equation}
1508: where the vector $Y^i$ is obtained by solving an elliptic equation
1509: of the form
1510: \begin{equation}\label{eq:Long_eqn}
1511: \LapLong Y^i = \nabla_jM^{ij}.
1512: \end{equation}
1513: The problem resides in the fact that the meaning of the TT component
1514: depends of the boundary conditions used in solving (\ref{eq:Long_eqn}).
1515:
1516: For the ConfTT/PhysTT cases we are actually solving for a vector $V^i$
1517: that is a linear combination of two components, one that solves an
1518: equation of the form of (\ref{eq:Long_eqn}) to obtain the TT component
1519: of $\tilde{M}^{ij}$ and one that solves the momentum constraint. But
1520: by imposing inner-boundary conditions of $V^i=0$, we don't specify the
1521: boundary conditions on either part independently. Nor is it clear
1522: what these boundary conditions should be. Since we cannot explicitly
1523: construct the TT component of the extrinsic curvature, we cannot
1524: eliminate it. Although it is not ideal, there is an alternative we
1525: can consider that does provide some insight into the importance of
1526: the initial choice of $\tilde{M}^{ij}$.
1527:
1528: %********************************
1529: \subsubsection{Black holes at rest}
1530: \label{sec:mConfTT:Rest}
1531: %********************************
1532:
1533: Consider the following numerical experiment for two black holes at
1534: rest: Given $\tilde M^{ij}$ from Eq.~(\ref{eq:BinaryKerrSchild-Mij}),
1535: make a transverse traceless decomposition by setting
1536: \begin{equation}\label{eq:mConfTT}
1537: 2N \tilde M^{ij}=\tilde M_{TT}^{ij}+(\tilde\Long Y)^{ij}
1538: \end{equation}
1539: where $\tilde\nabla_j \tilde M^{ij}_{TT}=0$ and $N=N_A+N_B-1$.
1540: Notice that we are decomposing $2N\tilde{M}^{ij}$, not $\tilde{M}^{ij}$.
1541: Taking
1542: the divergence of Eq.~(\ref{eq:mConfTT}) one finds
1543: \begin{equation}\label{eq:mConfTT-X}
1544: \tildeLapLong Y^i=\tilde\nabla_j\left(2N\tilde M^{ij}\right).
1545: \end{equation}
1546: The decomposition chosen in Eq.~(\ref{eq:mConfTT}) is motivated by the
1547: conformal thin sandwich decomposition. With this decomposition we can,
1548: in fact, use the shift vector $N^i$ to fix boundary conditions on
1549: $Y^i$, just as it was used to fix the boundary conditions in
1550: Eqs.~(\ref{eq:Sandwich-BC2}---\ref{eq:Sandwich-BC4}). For the black
1551: holes at rest in this case, we have $\Omega=0$. After solving
1552: Eq.~(\ref{eq:mConfTT-X}) for $Y^i$, we can construct a new
1553: conformal extrinsic curvature by
1554: \begin{equation}\label{eq:mConfTT-Mij}
1555: {\tilde{M}{'}}^{\;ij}=\frac{1}{2N}(\tilde\Long Y)^{ij}
1556: \end{equation}
1557: which is similar to what would result if we could eliminate $\tilde
1558: M^{ij}_{TT}$ from $\tilde M^{ij}$. Using $\tilde M^{\prime ij}$ in
1559: place of $\tilde M^{ij}$, we can again solve the conformal TT
1560: equations. The result of this modified conformal TT decomposition
1561: {\bf ``mConfTT''} is striking: Figure~\ref{fig:Compare-cuts} shows
1562: that mConfTT generates a conformal factor $\psi$ that is very similar
1563: to $\psi$ of CTS. mConfTT is also included in Table~\ref{tab:EADM}
1564: where it can be seen that the quantities $E_{ADM}/m$ and
1565: $E_{MPRC}/E_{ADM}$ differ only slightly between mConfTT and CTS.
1566:
1567: The fact that modification of the extrinsic curvature changes the
1568: ADM-energy by such a large amount underlines the importance of a
1569: careful choice for the extrinsic curvature $\tilde M^{ij}$ in
1570: ConfTT/PhysTT. The extremely good agreement between CTS and mConfTT
1571: is probably caused by our procedure to determine $\tilde M{'}^{ij}$.
1572: We force the extrinsic curvature of mConfTT into the form
1573: Eq.~(\ref{eq:mConfTT-Mij}). This is precisely the form of the
1574: extrinsic curvature in CTS, Eq.~(\ref{eq:SandwichTT-3}), even using
1575: the same function $N$ and the same boundary conditions on the vectors
1576: $Y^i$ and $\beta^i$.
1577:
1578: %****************************************************************
1579: \subsubsection{Black holes with angular momentum}
1580: %****************************************************************
1581:
1582: We now apply the modified conformal TT decomposition to the orbiting
1583: configurations of section~\ref{sec:resultOrbiting}. In the corotating
1584: frame, the black holes are at rest, and we start with
1585: $\tilde\gamma_{ij}$ and $\tilde M^{ij}$ of two black holes {\em at
1586: rest} with coordinate separation $d=11.860$. We now solve
1587: Eq.~(\ref{eq:mConfTT-X}) with
1588: \begin{equation}
1589: N=N_A+N_B-1
1590: \end{equation}
1591: and corotating boundary conditions on $Y^i$ [cf.\
1592: Eqs.~(\ref{eq:Sandwich-BC2})--(\ref{eq:Sandwich-BC4})]:
1593: \begin{subequations}
1594: \begin{align}
1595: Y^i&=N_{\!A}^i &&\mbox{sphere inside hole A,}\\
1596: Y^i&=N_B^i &&\mbox{sphere inside hole B,}\\
1597: Y^i&=\vec\Omega\times\vec r &&\mbox{outer boundary.}
1598: \end{align}
1599: \end{subequations}
1600: $N_{A/B}$ and $N_{A/B}^i$ are lapse and shift of individual Kerr-Schild
1601: black holes at rest. $\tilde M{'}^{ij}$ is again constructed by
1602: Eq.~(\ref{eq:mConfTT-Mij}) and used in solving the conformal TT equations.
1603:
1604: Results from this procedure are included in Table~\ref{tab:EADM2}.
1605: Again, mConfTT generates results very close to CTS. $E_{ADM}/m$
1606: changes by 1.8\% of the total mass between ConfTT and mConfTT, again
1607: highlighting the importance of the extrinsic curvature.
1608:
1609:
1610: %****************************************************************
1611: \subsection{Dependence on the size of the excised spheres}
1612: %****************************************************************
1613:
1614: The framework presented in this paper requires the excision of the
1615: singularities at the centers of each hole\footnote{Marronetti and
1616: Matzner\cite{Marronetti-Matzner:2000} effectively excised the
1617: centers, too, by using ``blending functions''.}. So far we have used
1618: $r_{exc}=2M$ or $r_{exc}=1.9M$ in order to impose boundary conditions
1619: close to the apparent horizons, but different choices can be made.
1620: Indeed, one might expect that the boundary conditions
1621: (\ref{eq:BC-ConfTT-PhysTT}) and (\ref{eq:BC-sandwich}) become ``better''
1622: farther inside the apparent horizon, where the metric and extrinsic
1623: curvature of that black hole dominate the superposed metric
1624: $\tilde\gamma_{ij}$ and superposed extrinsic curvature $\tilde
1625: M^{ij}$.
1626:
1627: In order to test this assumption, we solve the constraint equations
1628: for two black holes at rest for different radii $r_{exc}$. We find
1629: that for all three decompositions, the data sets depend strongly on
1630: the radius of the excised spheres.
1631:
1632: \begin{figure}
1633: \centering
1634: % Plot generated by Neville/DataEric/Run/SP2/PlotCuts.sm
1635: \includegraphics[scale=0.38]{Fig9.eps}
1636: \caption{Plots of $\psi$ and $V^x$ along the positive $x$-axis
1637: for ConfTT for different radii $r_{exc}=2M, M, 0.5M, 0.2M$. The
1638: excised spheres are centered on the $x$-axis at $x=\pm 5$. The
1639: position where a line terminates gives $r_{exc}$ for that line.}
1640: \label{fig:ConfTT-cuts}
1641: \end{figure}
1642:
1643: \begin{figure}
1644: \centerline{\includegraphics[scale=0.38]{Fig10.eps}}
1645: \caption{\label{fig:ConfTT-AH}Apparent horizons for ConfTT
1646: with different radii of excised spheres. Results shown are for
1647: $r_{exc}=2M$ (long dashed line), $M$ (dotted line), $0.5M$ (short
1648: dashed line) and $0.2M$ (outer solid line). The inner solid line
1649: is a circle with radius 2. The insert shows a parametric plot of
1650: $r(\phi)-2$, which emphasizes the differences between the
1651: different apparent horizons.}
1652: \end{figure}
1653:
1654:
1655: \begin{table}
1656: \caption{\label{tab:TT}
1657: Solutions of solving ConfTT for different radii of
1658: the excised spheres, $r_{exc}$. The results for PhysTT
1659: are nearly identical.
1660: }
1661: \begin{ruledtabular}
1662: \begin{tabular}{cccccc}
1663: $r_{exc}$ & $E_{ADM}$ & $A_{AH}$ & $\ell$ & $E_{ADM}/m$ & $\ell/E_{ADM}$
1664: \rule[-.65em]{0.em}{1.3em}\\\hline
1665: \multicolumn{6}{c}{ Conformal TT }\\
1666: 2.0 & 2.0649 & 57.737 & 8.062 & 0.9633 & 3.904 \\
1667: 1.0 & 2.0682 & 57.825 & 8.101 & 0.9641 & 3.917 \\
1668: 0.5 & 2.0808 & 58.520 & 8.101 & 0.9642 & 3.893 \\
1669: 0.2 & 2.0978 & 59.514 & 8.093 & 0.9640 & 3.858 \\
1670: 0.1 & 2.1064 & 60.025 & 8.089 & 0.9638 & 3.840 \\
1671: \end{tabular}
1672: \end{ruledtabular}
1673: \end{table}
1674:
1675: Figure~\ref{fig:ConfTT-cuts} presents plots of the conformal factor
1676: $\psi$ and $V^x$ for ConfTT with different $r_{exc}$. There is no
1677: clear sign of convergence of $\psi$ as $r_{exc}\to 0$. For
1678: $r_{exc}=0.2M$, the conformal factor $\psi$ even oscillates close to
1679: the excised sphere. Table~\ref{tab:TT} displays various
1680: quantities for the ConfTT decomposition for
1681: different $r_{exc}$.
1682: As $r_{exc}$ varies
1683: between $2.0M$ and $0.1M$, the ADM-energy varies between $2.065$ and
1684: $2.106$, whereas the apparent horizon area changes by nearly 4\%. The
1685: apparent horizons move around somewhat as $r_{exc}$ changes.
1686: Figure~\ref{fig:ConfTT-AH} shows the location of the apparent horizons for
1687: different $r_{exc}$.
1688:
1689: \begin{figure}
1690: \centering
1691: %Plot generated by Neville/DataEric/Run/SP2/PlotCuts.sm
1692: \includegraphics[scale=0.38]{Fig11.eps}
1693: \caption{\label{fig:Sandwich-cuts-N1N2}
1694: Cuts through $\psi$ and $\beta^x$ for CTS-mult for different radii
1695: $r_{exc}$. Here $\tilde \alpha=N_{\!A}N_B$ and the boundary
1696: condition on $\psi$ at the excised spheres is $d\psi/dr=0$. The
1697: curves for $\beta^x$ are shifted up by 0.5 for $x<5$, and are
1698: shifted down by 0.5 for $x>5$ to allow for better plotting.
1699: $d\psi/dr$ approaches zero at the inner boundary on scales too
1700: small to be seen in this figure.
1701: }
1702: \end{figure}
1703:
1704:
1705: For CTS-add (with $\tilde\alpha=N_A+N_B-1$), the initial-data sets
1706: seem to diverge as $r_{exc}$ is decreased. This has to be expected,
1707: since this choice for $\tilde\alpha$ changes sign if the excised
1708: spheres become sufficiently small. Changing to $\tilde\alpha=N_AN_B$
1709: so that the lapse does not change sign reduces this divergent behavior. Von
1710: Neumann boundary conditions on $\psi$ at the excised spheres,
1711: \begin{equation}
1712: \frac{\partial\psi}{\partial r}=0,
1713: \end{equation}
1714: lead to an increase in $A_{AH}$ especially for large excised spheres.
1715: This combination of lapse $\tilde\alpha$ and boundary conditions
1716: exhibits the smallest variations in $E_{ADM}/m$; cuts through $\psi$,
1717: $\beta^x$ and through the apparent horizons are shown in
1718: Figs.~\ref{fig:Sandwich-cuts-N1N2} and \ref{fig:Sandwich-AH-2}. From
1719: the three examined combinations of lapse and boundary conditions, the
1720: one shown behaves best, but there is still no convincing sign of
1721: convergence.
1722:
1723: Table~\ref{tab:CTS} presents ADM-energies and apparent
1724: horizon areas and masses for CTS with different $r_{exc}$ and
1725: different choices of lapse and boundary condition. From the unscaled
1726: ADM-energy $E_{ADM}$ it is apparent that $\tilde\alpha=N_A+N_B-1$
1727: diverges most strongly. Note that between all choices of lapse,
1728: boundary conditions and $r_{exc}$, the unscaled quantities $E_{ADM}$,
1729: $M_{AH}$, and $\ell$ exhibit a much broader variation than the scaled
1730: quantities $E_{ADM}/m$ and $\ell/E_{ADM}$.
1731:
1732: \begin{figure}
1733: \centerline{\includegraphics[scale=0.38]{Fig12.eps}}
1734: \caption{\label{fig:Sandwich-AH-2}
1735: Apparent horizons for CTS with $\tilde\alpha=N_A\,N_B$ and inner
1736: boundary condition $d\psi/dr=0$. The different curves belong to
1737: different $r_{exc}$ as explained in Fig.~\ref{fig:ConfTT-AH} }
1738: \end{figure}
1739:
1740: \begin{table}
1741: \caption{\label{tab:CTS}
1742: Solutions of CTS as a function of radius of excised
1743: spheres, $r_{exc}$. Different choices of the lapse $\tilde\alpha$ and
1744: boundary conditions for $\psi$ at the excised spheres are explored.
1745: }
1746: \begin{ruledtabular}
1747: \begin{tabular}{ccrccc}
1748: $r_{exc}$ & $E_{ADM}$ & $A_{AH}\;$& $\ell$ & $E_{ADM}/m$ & $\ell/E_{ADM}$
1749: \rule[-.5em]{0.em}{1.5em}\\\hline
1750: \multicolumn{6}{c}{ $\tilde\alpha=N_A+N_B-1,\;\; \psi=1$ }\\
1751: 2.0 & 2.0812 & 62.312 & 8.039 & 0.9346 & 3.863 \\
1752: 1.0 & 2.1846 & 68.279 & 8.000 & 0.9372 & 3.662 \\
1753: 0.5 & 2.3085 & 76.253 & 7.925 & 0.9371 & 3.433 \\
1754: 0.2 & 2.5463 & 93.534 & 7.750 & 0.9333 & 3.044 \\
1755: 0.1 & 2.8543 & 118.834 & 7.489 & 0.9282 & 2.624 \\
1756: \hline
1757: \multicolumn{6}{c}{ $\tilde\alpha=N_AN_B,\;\; \psi=1$ }\\
1758: 2.0 & 2.0585 & 60.811 & 8.080 & 0.9358 & 3.925 \\
1759: 1.0 & 2.1216 & 64.080 & 8.044 & 0.9395 & 3.792 \\
1760: 0.5 & 2.1696 & 66.790 & 8.017 & 0.9411 & 3.695 \\
1761: 0.2 & 2.2120 & 69.456 & 7.991 & 0.9409 & 3.613 \\
1762: 0.1 & 2.2326 & 70.809 & 7.978 & 0.9405 & 3.573 \\
1763: \hline
1764: \multicolumn{6}{c}{ $\tilde\alpha=N_AN_B,\;\; \partial\psi/\partial r=0$ }\\
1765: 2.0 & 2.1110 & 64.229 & 8.085 & 0.9337 & 3.830 \\
1766: 1.0 & 2.1533 & 66.128 & 8.030 & 0.9387 & 3.729 \\
1767: 0.5 & 2.1794 & 67.427 & 8.011 & 0.9409 & 3.676 \\
1768: 0.2 & 2.2136 & 69.559 & 7.990 & 0.9409 & 3.609 \\
1769: 0.1 & 2.2330 & 70.836 & 7.978 & 0.9405 & 3.573 \\
1770: \end{tabular}
1771: \end{ruledtabular}
1772: \end{table}
1773:
1774: %****************************************************************
1775: \section{Discussion}
1776: \label{sec:Discussion}
1777: %****************************************************************
1778:
1779: Our results clearly show that different decompositions lead to
1780: different initial-data sets, even when seemingly similar
1781: choices for the freely specifiable pieces are used. From
1782: Tables~\ref{tab:EADM} and \ref{tab:EADM2}, one sees that $E_{ADM}/m$
1783: changes by as much as 0.029 between ConfTT/PhysTT and CTS. The
1784: difference between ConfTT/PhysTT and the inversion symmetric data is
1785: even larger, 0.047. These numbers seem to be small; however, current
1786: evolutions of binary data usually find the total energy emitted in
1787: gravitational radiation $E_{GW}/m$ to be between 0.01 and 0.03
1788: \cite{Baker-Bruegmann-etal:2001,Alcubierre:2001,Brandt-Corell-etal:2000},
1789: which is the same order of magnitude as the changes in $E_{ADM}/m$ we
1790: find. This means that, in principle, most of the energy radiated in
1791: these simulations could originate from ``spurious'' energy in the
1792: system and not from the dynamics of the binary system we are
1793: interested in.
1794:
1795: These findings highlight the fact that current binary black hole
1796: initial data sets are inadequate for the task of accurately describing
1797: realistic binary systems. We see that the choices of the conformal
1798: 3-geometry $\tilde\gamma_{ij}$ and the freely specifiable portions of
1799: the extrinsic curvature, embedded in $\tilde{M}^{ij}$, influence the
1800: content of the initial data at a significant level. Furthermore, the
1801: results suggest that small changes in the free data associated with
1802: the extrinsic curvature are more significant than small changes in the
1803: choice of $\tilde\gamma_{ij}$.\footnote{Following submission of this
1804: paper, a preprint by Damour et
1805: al.\cite{Damour-Gourgoulhon-Grandclement:grqc0204011} has appeared
1806: that lends support to our idea that the extrinsic curvature plays a
1807: key role in constructing quasi-equilibrium binary black hole initial
1808: data.} This assertion is supported by the fact that $E_{ADM}/m$
1809: is consistently larger for the ConfTT solutions than for the CTS
1810: solutions but the two approaches can be made to produce quite
1811: consistent results by using the modified extrinsic curvature of the
1812: mConfTT method. All of these decompositions use the same non-flat
1813: conformal metric, but differ in the extrinsic curvature. On the other
1814: hand, results for the conformally-flat inversion-symmetric data agree
1815: rather well with the results from the CTS method when we consider
1816: orbiting black holes. For black holes at rest, CTS differs from the
1817: inversion symmetric data, which seems to contradict our conclusion.
1818: However, this difference is likely due to the time-symmetric nature of
1819: the inversion symmetric data, which is especially adapted to the
1820: time-symmetry of the particular configuration of ``two black holes at
1821: rest''.
1822:
1823: Improved binary black hole initial data will require choices for the
1824: freely specifiable data that are physically motivated, rather than
1825: chosen for computational convenience. The same is true for the
1826: boundary conditions used in solving the constraints. The boundary
1827: conditions used in this paper carry the implicit assumptions that the
1828: approximate metric and extrinsic curvature are correct at the
1829: excision boundaries and that the value of the single-hole Kerr-Schild
1830: shift at the excision boundary is correct in a multi-hole situation.
1831: This is clearly not true, but we might hope that the impact of the
1832: error in this choice would diminish as we decrease the radius of the
1833: excision boundary. However, our results presented in
1834: Tables~\ref{tab:TT} and \ref{tab:CTS} do not support this conjecture.
1835: Examining the change in $E_{ADM}/m$ as we vary $r_{exc}$ shows only a
1836: small change, but more importantly, it shows no sign of converging as
1837: we decrease $r_{exc}$. The effects of changing $r_{exc}$ are much
1838: more significant for $\ell/E_{ADM}$, changing its value by as much as
1839: 10\% in the case of CTS-mult for the range of values considered.
1840: Furthermore, as with the energy, we see no sign of convergence in
1841: $\ell/E_{ADM}$ as $r_{exc}$ decreases. Interestingly, although the
1842: solutions show no sign of convergence as we shrink the excision
1843: radius, we do find that the dimensionless quantities $E_{ADM}/m$ and
1844: $\ell/E_{ADM}$ do become independent of the choice of the
1845: inner-boundary condition as $r_{exc}$ decreases. This can be seen in
1846: comparing the result in Table~\ref{tab:CTS} for the cases using
1847: $\psi=1$ and $\partial\psi/\partial r=0$ as inner-boundary conditions.
1848: Additional tests, not reported in this paper, further support this
1849: assertion.
1850:
1851:
1852:
1853: %****************************************************************
1854: \section{Conclusion}
1855: %****************************************************************
1856:
1857: Using a new elliptic solver capable of solving the initial-value
1858: problem of general relativity for any of three different
1859: decompositions and any choice for the freely specifiable data, we have
1860: examined data sets representing binary black hole spacetimes. We find
1861: that the choices for the freely specifiable data currently in use are
1862: inadequate for the task of simulating the gravitational radiation
1863: produced in astrophysically realistic situations. In particular,
1864: we studied the results of using a superposition of two Kerr-Schild
1865: black holes to fix the freely specifiable data and compared them
1866: to the results obtained from conformally flat initial data.
1867:
1868: Although the new Kerr-Schild based data provide a valuable point of
1869: comparison, it is not clear that the data produced are significantly
1870: superior to previous conformally-flat data. What is clear is that the
1871: choice of the freely specifiable data will be very important in
1872: constructing astrophysically realistic binary black hole initial data.
1873: Progress will require that these data, {\em and} the boundary
1874: conditions needed to solve the constraints, must be chosen based on
1875: physical grounds rather than computational convenience.
1876:
1877: How can better initial data be achieved and how can the quality of
1878: initial data be measured? We believe that the conformal thin sandwich
1879: decomposition will be especially useful. Genuine radiative degrees of
1880: freedom cannot {\em in principle} be recognized on a single time
1881: slice. The conformal thin sandwich method uses in effect two nearby
1882: surfaces, giving it a potential advantage over other methods. Also,
1883: it avoids much of the uncertainty related to specifying a conformal
1884: extrinsic curvature.
1885: Moreover, the conformal thin sandwich approach is especially well
1886: suited for the most interesting configurations, a black-hole binary in
1887: a quasi-equilibrium orbit. In this case time derivatives of all
1888: quantities are small and the choice $\tilde u^{ij}=0$ is physically
1889: motivated. One should exploit the condition of quasi-equilibrium as
1890: fully as possible, i.e. one should use the conformal thin sandwich
1891: approach together with the constant $K$ equation, $\partial_tK=0$. The
1892: latter yields another elliptic equation for the lapse which removes
1893: the arbitrariness inherent in choosing a conformal lapse
1894: $\tilde\alpha$. One will also need more physical boundary conditions.
1895: Work in this direction was begun
1896: in\cite{Gourgoulhon-Grandclement-Bonazzola:2001a,
1897: Grandclement-Gourgoulhon-Bonazzola:2001b} and refined in
1898: \cite{Cook:2001}.
1899:
1900: Ultimately, the gravitational wave content of an initial-data set can
1901: be determined only by long term evolutions. One must compute an
1902: initial-data set representing a binary black hole in quasi-circular
1903: orbit and evolve it. Then one must repeat this process with an
1904: initial-data set representing the {\em same} binary black hole, say, one
1905: orbital period earlier, and evolve that data set, too. If both
1906: evolutions lead to the same gravitational waves (modulo time offset)
1907: then one can be confident that the gravitational radiation is indeed
1908: astrophysically realistic. This approach has recently been used for
1909: the first time in conjunction with conformally flat puncture data,
1910: where it proved remarkably successful
1911: \cite{Baker-Campanelli-etal:astroph0202469}.
1912:
1913:
1914: \acknowledgments We thank Lawrence Kidder, Mark Scheel, and James
1915: York for helpful discussions. This work was supported in part by NSF
1916: grants PHY-9800737 and PHY-9900672 to Cornell University, and by NSF
1917: grant PHY-9988581 to Wake Forest University. Computations were
1918: performed on the IBM SP2 of the Department of Physics, Wake Forest
1919: University, with support from an IBM SUR grant.
1920:
1921:
1922: %\bibliography{Definitions,ID}
1923: \begin{thebibliography}{40}
1924: \expandafter\ifx\csname natexlab\endcsname\relax\def\natexlab#1{#1}\fi
1925: \expandafter\ifx\csname bibnamefont\endcsname\relax
1926: \def\bibnamefont#1{#1}\fi
1927: \expandafter\ifx\csname bibfnamefont\endcsname\relax
1928: \def\bibfnamefont#1{#1}\fi
1929: \expandafter\ifx\csname citenamefont\endcsname\relax
1930: \def\citenamefont#1{#1}\fi
1931: \expandafter\ifx\csname url\endcsname\relax
1932: \def\url#1{\texttt{#1}}\fi
1933: \expandafter\ifx\csname urlprefix\endcsname\relax\def\urlprefix{URL }\fi
1934: \providecommand{\bibinfo}[2]{#2}
1935: \providecommand{\eprint}[2][]{\url{#2}}
1936:
1937: \bibitem[{\citenamefont{Baker et~al.}(2001)\citenamefont{Baker, Br{\"u}gmann,
1938: Campanelli, Lousto, and Takahashi}}]{Baker-Bruegmann-etal:2001}
1939: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Baker}},
1940: \bibinfo{author}{\bibfnamefont{B.}~\bibnamefont{Br{\"u}gmann}},
1941: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Campanelli}},
1942: \bibinfo{author}{\bibfnamefont{C.~O.} \bibnamefont{Lousto}},
1943: \bibnamefont{and}
1944: \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Takahashi}},
1945: \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{87}},
1946: \bibinfo{pages}{121103} (\bibinfo{year}{2001}).
1947:
1948: \bibitem[{\citenamefont{Alcubierre et~al.}(2001)\citenamefont{Alcubierre,
1949: Benger, Br{\"u}gmann, Lanfermann, Nerger, Seidel, and
1950: Takahashi}}]{Alcubierre:2001}
1951: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Alcubierre}},
1952: \bibinfo{author}{\bibfnamefont{W.}~\bibnamefont{Benger}},
1953: \bibinfo{author}{\bibfnamefont{B.}~\bibnamefont{Br{\"u}gmann}},
1954: \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Lanfermann}},
1955: \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Nerger}},
1956: \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Seidel}}, \bibnamefont{and}
1957: \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Takahashi}},
1958: \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{87}},
1959: \bibinfo{pages}{271103} (\bibinfo{year}{2001}).
1960:
1961: \bibitem[{\citenamefont{Baker et~al.}(2002)\citenamefont{Baker, Campanelli,
1962: Lousto, and Takahashi}}]{Baker-Campanelli-etal:astroph0202469}
1963: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Baker}},
1964: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Campanelli}},
1965: \bibinfo{author}{\bibfnamefont{C.~O.} \bibnamefont{Lousto}},
1966: \bibnamefont{and}
1967: \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Takahashi}},
1968: \bibinfo{journal}{astro-ph/0202469} (\bibinfo{year}{2002}).
1969:
1970: \bibitem[{\citenamefont{York{,}~Jr.}(1979)}]{York:1979}
1971: \bibinfo{author}{\bibfnamefont{J.~W.} \bibnamefont{York{,}~Jr.}}, in
1972: \emph{\bibinfo{booktitle}{Sources of Gravitational Radiation}}, edited by
1973: \bibinfo{editor}{\bibfnamefont{L.~L.} \bibnamefont{Smarr}}
1974: (\bibinfo{address}{Cambridge University Press, Cambridge, UK},
1975: \bibinfo{year}{1979}), pp. \bibinfo{pages}{83--126}.
1976:
1977: \bibitem[{\citenamefont{Murchadha and
1978: York{,}~Jr.}(1974{\natexlab{a}})}]{Murchadha-York:1974a}
1979: \bibinfo{author}{\bibfnamefont{N.~{\'O}.} \bibnamefont{Murchadha}}
1980: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{J.~W.}
1981: \bibnamefont{York{,}~Jr.}}, \bibinfo{journal}{Phys. Rev. D}
1982: \textbf{\bibinfo{volume}{10}}, \bibinfo{pages}{428}
1983: (\bibinfo{year}{1974}{\natexlab{a}}).
1984:
1985: \bibitem[{\citenamefont{Murchadha and
1986: York{,}~Jr.}(1974{\natexlab{b}})}]{Murchadha-York:1974b}
1987: \bibinfo{author}{\bibfnamefont{N.~{\'O}.} \bibnamefont{Murchadha}}
1988: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{J.~W.}
1989: \bibnamefont{York{,}~Jr.}}, \bibinfo{journal}{Phys. Rev. D}
1990: \textbf{\bibinfo{volume}{10}}, \bibinfo{pages}{437}
1991: (\bibinfo{year}{1974}{\natexlab{b}}).
1992:
1993: \bibitem[{\citenamefont{Murchadha and York{,}~Jr.}(1976)}]{Murchadha-York:1976}
1994: \bibinfo{author}{\bibfnamefont{N.~{\'O}.} \bibnamefont{Murchadha}}
1995: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{J.~W.}
1996: \bibnamefont{York{,}~Jr.}}, \bibinfo{journal}{Gen. Relativ. Gravit.}
1997: \textbf{\bibinfo{volume}{7}}, \bibinfo{pages}{257} (\bibinfo{year}{1976}).
1998:
1999: \bibitem[{\citenamefont{York{,}~Jr.}(1999)}]{York:1999}
2000: \bibinfo{author}{\bibfnamefont{J.~W.} \bibnamefont{York{,}~Jr.}},
2001: \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{82}},
2002: \bibinfo{pages}{1350} (\bibinfo{year}{1999}).
2003:
2004: \bibitem[{\citenamefont{Bowen}(1979)}]{Bowen:1979}
2005: \bibinfo{author}{\bibfnamefont{J.~M.} \bibnamefont{Bowen}},
2006: \bibinfo{journal}{Gen. Relativ. Gravit.} \textbf{\bibinfo{volume}{11}},
2007: \bibinfo{pages}{227} (\bibinfo{year}{1979}).
2008:
2009: \bibitem[{\citenamefont{Bowen and York{,}~Jr.}(1980)}]{Bowen-York:1980}
2010: \bibinfo{author}{\bibfnamefont{J.~M.} \bibnamefont{Bowen}} \bibnamefont{and}
2011: \bibinfo{author}{\bibfnamefont{J.~W.} \bibnamefont{York{,}~Jr.}},
2012: \bibinfo{journal}{Phys. Rev. D} \textbf{\bibinfo{volume}{21}},
2013: \bibinfo{pages}{2047} (\bibinfo{year}{1980}).
2014:
2015: \bibitem[{\citenamefont{Kulkarni et~al.}(1983)\citenamefont{Kulkarni, Shepley,
2016: and York{,}~Jr.}}]{Kulkarni-Shepley-York:1983}
2017: \bibinfo{author}{\bibfnamefont{A.~D.} \bibnamefont{Kulkarni}},
2018: \bibinfo{author}{\bibfnamefont{L.~C.} \bibnamefont{Shepley}},
2019: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{J.~W.}
2020: \bibnamefont{York{,}~Jr.}}, \bibinfo{journal}{Phys. Lett. A}
2021: \textbf{\bibinfo{volume}{96A}}, \bibinfo{pages}{228} (\bibinfo{year}{1983}).
2022:
2023: \bibitem[{\citenamefont{Thornburg}(1987)}]{Thornburg:1987}
2024: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Thornburg}},
2025: \bibinfo{journal}{Class. Quantum Gravit.} \textbf{\bibinfo{volume}{4}},
2026: \bibinfo{pages}{1119} (\bibinfo{year}{1987}).
2027:
2028: \bibitem[{\citenamefont{Cook et~al.}(1993)\citenamefont{Cook, Choptuik, Dubal,
2029: Klasky, Matzner, and Oliveira}}]{Cook-Choptuik-etal:1993}
2030: \bibinfo{author}{\bibfnamefont{G.~B.} \bibnamefont{Cook}},
2031: \bibinfo{author}{\bibfnamefont{M.~W.} \bibnamefont{Choptuik}},
2032: \bibinfo{author}{\bibfnamefont{M.~R.} \bibnamefont{Dubal}},
2033: \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Klasky}},
2034: \bibinfo{author}{\bibfnamefont{R.~A.} \bibnamefont{Matzner}},
2035: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{S.~R.}
2036: \bibnamefont{Oliveira}}, \bibinfo{journal}{Phys. Rev. D}
2037: \textbf{\bibinfo{volume}{47}}, \bibinfo{pages}{1471} (\bibinfo{year}{1993}).
2038:
2039: \bibitem[{\citenamefont{Brandt and Br{\"u}gmann}(1997)}]{Brandt-Brugmann:1997}
2040: \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Brandt}} \bibnamefont{and}
2041: \bibinfo{author}{\bibfnamefont{B.}~\bibnamefont{Br{\"u}gmann}},
2042: \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{78}},
2043: \bibinfo{pages}{3606} (\bibinfo{year}{1997}).
2044:
2045: \bibitem[{\citenamefont{Rieth}(1997)}]{Rieth:1997}
2046: \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Rieth}}, in
2047: \emph{\bibinfo{booktitle}{Mathematics of Gravitation. Part II. Gravitational
2048: Wave Detection}}, edited by
2049: \bibinfo{editor}{\bibfnamefont{A.}~\bibnamefont{Kr{\'o}lak}}
2050: (\bibinfo{publisher}{Polish Academy of Sciences, Institute of Mathematics,
2051: Warsaw}, \bibinfo{year}{1997}), pp. \bibinfo{pages}{71--74}.
2052:
2053: \bibitem[{\citenamefont{Damour et~al.}(2000)\citenamefont{Damour, Jaranowski,
2054: and Sch{\"a}fer}}]{Damour-Jaranowski-Schaefer:2000b}
2055: \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Damour}},
2056: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Jaranowski}},
2057: \bibnamefont{and}
2058: \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Sch{\"a}fer}},
2059: \bibinfo{journal}{Phys. Rev. D} \textbf{\bibinfo{volume}{62}},
2060: \bibinfo{pages}{084011} (\bibinfo{year}{2000}).
2061:
2062: \bibitem[{\citenamefont{Monroe}(1976)}]{Monroe:1976}
2063: \bibinfo{author}{\bibfnamefont{D.~K.} \bibnamefont{Monroe}}, Ph.D. thesis,
2064: \bibinfo{school}{University of North Carolina} (\bibinfo{year}{1976}).
2065:
2066: \bibitem[{\citenamefont{Garat and Price}(2000)}]{Garat-Price:2000}
2067: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Garat}} \bibnamefont{and}
2068: \bibinfo{author}{\bibfnamefont{R.~H.} \bibnamefont{Price}},
2069: \bibinfo{journal}{Phys. Rev. D} \textbf{\bibinfo{volume}{61}},
2070: \bibinfo{pages}{124011} (\bibinfo{year}{2000}).
2071:
2072: \bibitem[{\citenamefont{Pfeiffer et~al.}(2000)\citenamefont{Pfeiffer,
2073: Teukolsky, and Cook}}]{Pfeiffer-Teukolsky-Cook:2000}
2074: \bibinfo{author}{\bibfnamefont{H.~P.} \bibnamefont{Pfeiffer}},
2075: \bibinfo{author}{\bibfnamefont{S.~A.} \bibnamefont{Teukolsky}},
2076: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{G.~B.} \bibnamefont{Cook}},
2077: \bibinfo{journal}{Phys. Rev. D} \textbf{\bibinfo{volume}{62}},
2078: \bibinfo{pages}{104018} (\bibinfo{year}{2000}).
2079:
2080: \bibitem[{\citenamefont{Lousto and Price}(1998)}]{Lousto-Price:1998}
2081: \bibinfo{author}{\bibfnamefont{C.~O.} \bibnamefont{Lousto}} \bibnamefont{and}
2082: \bibinfo{author}{\bibfnamefont{R.~H.} \bibnamefont{Price}},
2083: \bibinfo{journal}{Phys. Rev. D} \textbf{\bibinfo{volume}{57}},
2084: \bibinfo{pages}{1073} (\bibinfo{year}{1998}).
2085:
2086: \bibitem[{\citenamefont{Matzner et~al.}(1999)\citenamefont{Matzner, Huq, and
2087: Shoemaker}}]{Matzner-Huq-Shoemaker:1999}
2088: \bibinfo{author}{\bibfnamefont{R.~A.} \bibnamefont{Matzner}},
2089: \bibinfo{author}{\bibfnamefont{M.~F.} \bibnamefont{Huq}}, \bibnamefont{and}
2090: \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Shoemaker}},
2091: \bibinfo{journal}{Phys. Rev. D} \textbf{\bibinfo{volume}{59}},
2092: \bibinfo{pages}{024015} (\bibinfo{year}{1999}).
2093:
2094: \bibitem[{\citenamefont{Marronetti and
2095: Matzner}(2000)}]{Marronetti-Matzner:2000}
2096: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Marronetti}} \bibnamefont{and}
2097: \bibinfo{author}{\bibfnamefont{R.~A.} \bibnamefont{Matzner}},
2098: \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{85}},
2099: \bibinfo{pages}{5500} (\bibinfo{year}{2000}).
2100:
2101: \bibitem[{\citenamefont{Gourgoulhon et~al.}(2002)\citenamefont{Gourgoulhon,
2102: Grandcl{\'e}ment, and Bonazzola}}]{Gourgoulhon-Grandclement-Bonazzola:2001a}
2103: \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Gourgoulhon}},
2104: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Grandcl{\'e}ment}},
2105: \bibnamefont{and}
2106: \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Bonazzola}},
2107: \bibinfo{journal}{Phys. Rev. D} \textbf{\bibinfo{volume}{65}},
2108: \bibinfo{pages}{044020} (\bibinfo{year}{2002}).
2109:
2110: \bibitem[{\citenamefont{Grandcl{\'e}ment
2111: et~al.}(2002)\citenamefont{Grandcl{\'e}ment, Gourgoulhon, and
2112: Bonazzola}}]{Grandclement-Gourgoulhon-Bonazzola:2001b}
2113: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Grandcl{\'e}ment}},
2114: \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Gourgoulhon}},
2115: \bibnamefont{and}
2116: \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Bonazzola}},
2117: \bibinfo{journal}{Phys. Rev. D} \textbf{\bibinfo{volume}{65}},
2118: \bibinfo{pages}{044021} (\bibinfo{year}{2002}).
2119:
2120: \bibitem[{\citenamefont{Bonazzola et~al.}(1993)\citenamefont{Bonazzola,
2121: Gourgoulhon, Salgado, and Marck}}]{Bonazzola-Gourgoulhon-etal:1993}
2122: \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Bonazzola}},
2123: \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Gourgoulhon}},
2124: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Salgado}}, \bibnamefont{and}
2125: \bibinfo{author}{\bibfnamefont{J.-A.} \bibnamefont{Marck}},
2126: \bibinfo{journal}{Astron. \& Astrophys.} \textbf{\bibinfo{volume}{278}},
2127: \bibinfo{pages}{421} (\bibinfo{year}{1993}).
2128:
2129: \bibitem[{\citenamefont{Bonazzola et~al.}(1998)\citenamefont{Bonazzola,
2130: Gourgoulhon, and Marck}}]{Bonazzola-Gourgoulhon-Marck:1998}
2131: \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Bonazzola}},
2132: \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Gourgoulhon}},
2133: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{J.-A.} \bibnamefont{Marck}},
2134: \bibinfo{journal}{Phys. Rev. D} \textbf{\bibinfo{volume}{58}},
2135: \bibinfo{pages}{104020} (\bibinfo{year}{1998}).
2136:
2137: \bibitem[{\citenamefont{Kidder et~al.}(2000)\citenamefont{Kidder, Scheel,
2138: Teukolsky, Carlson, and Cook}}]{Kidder-Scheel-etal:2000}
2139: \bibinfo{author}{\bibfnamefont{L.~E.} \bibnamefont{Kidder}},
2140: \bibinfo{author}{\bibfnamefont{M.~A.} \bibnamefont{Scheel}},
2141: \bibinfo{author}{\bibfnamefont{S.~A.} \bibnamefont{Teukolsky}},
2142: \bibinfo{author}{\bibfnamefont{E.~D.} \bibnamefont{Carlson}},
2143: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{G.~B.} \bibnamefont{Cook}},
2144: \bibinfo{journal}{Phys. Rev. D} \textbf{\bibinfo{volume}{62}},
2145: \bibinfo{pages}{084032} (\bibinfo{year}{2000}).
2146:
2147: \bibitem[{\citenamefont{Gourgoulhon et~al.}(2001)\citenamefont{Gourgoulhon,
2148: Grandcl{\'e}ment, Taniguchi, Marck, and
2149: Bonazzola}}]{Gourgoulhon-Grandclement-etal:2001}
2150: \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Gourgoulhon}},
2151: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Grandcl{\'e}ment}},
2152: \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Taniguchi}},
2153: \bibinfo{author}{\bibfnamefont{J.-A.} \bibnamefont{Marck}}, \bibnamefont{and}
2154: \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Bonazzola}},
2155: \bibinfo{journal}{Phys. Rev. D} \textbf{\bibinfo{volume}{63}},
2156: \bibinfo{pages}{064029} (\bibinfo{year}{2001}).
2157:
2158: \bibitem[{\citenamefont{Kidder et~al.}(2001)\citenamefont{Kidder, Scheel, and
2159: Teukolsky}}]{Kidder-Scheel-Teukolsky:2001}
2160: \bibinfo{author}{\bibfnamefont{L.~E.} \bibnamefont{Kidder}},
2161: \bibinfo{author}{\bibfnamefont{M.~A.} \bibnamefont{Scheel}},
2162: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{S.~A.}
2163: \bibnamefont{Teukolsky}}, \bibinfo{journal}{Phys. Rev. D}
2164: \textbf{\bibinfo{volume}{64}}, \bibinfo{pages}{064017}
2165: (\bibinfo{year}{2001}).
2166:
2167: \bibitem[{\citenamefont{Ansorg et~al.}(2002)\citenamefont{Ansorg,
2168: Kleinw{\"a}chter, and Meinel}}]{Ansorg-Kleinwaechter-Meinel:2002}
2169: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Ansorg}},
2170: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Kleinw{\"a}chter}},
2171: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Meinel}},
2172: \bibinfo{journal}{Astron. \& Astrophys.} \textbf{\bibinfo{volume}{381}},
2173: \bibinfo{pages}{L49} (\bibinfo{year}{2002}).
2174:
2175: \bibitem[{\citenamefont{Pfeiffer et~al.}(2002)\citenamefont{Pfeiffer, Kidder,
2176: Scheel, and Teukolsky}}]{Pfeiffer-Kidder-etal:2002}
2177: \bibinfo{author}{\bibfnamefont{H.~P.} \bibnamefont{Pfeiffer}},
2178: \bibinfo{author}{\bibfnamefont{L.~E.} \bibnamefont{Kidder}},
2179: \bibinfo{author}{\bibfnamefont{M.~A.} \bibnamefont{Scheel}},
2180: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{S.~A.}
2181: \bibnamefont{Teukolsky}} (\bibinfo{year}{2002}),
2182: \bibinfo{note}{gr-qc/0202096}.
2183:
2184: \bibitem[{\citenamefont{Cook}(2000)}]{Cook:2000}
2185: \bibinfo{author}{\bibfnamefont{G.~B.} \bibnamefont{Cook}},
2186: \bibinfo{journal}{Living Rev. Relativity} \textbf{\bibinfo{volume}{3}}
2187: (\bibinfo{year}{2000}), \bibinfo{note}{[Online Article]: cited on Aug 11,
2188: 2001},
2189: \urlprefix\url{http://www.livingreviews.org/Articles/Volume3/2000-5cook/}.
2190:
2191: \bibitem[{\citenamefont{Misner}(1963)}]{Misner:1963}
2192: \bibinfo{author}{\bibfnamefont{C.~W.} \bibnamefont{Misner}},
2193: \bibinfo{journal}{Annals of Physics} \textbf{\bibinfo{volume}{24}},
2194: \bibinfo{pages}{102} (\bibinfo{year}{1963}).
2195:
2196: \bibitem[{\citenamefont{Cook}(1991)}]{Cook:1991}
2197: \bibinfo{author}{\bibfnamefont{G.~B.} \bibnamefont{Cook}},
2198: \bibinfo{journal}{Phys. Rev. D} \textbf{\bibinfo{volume}{44}},
2199: \bibinfo{pages}{2983} (\bibinfo{year}{1991}).
2200:
2201: \bibitem[{\citenamefont{York{,}~Jr.}(2002)}]{York:2002}
2202: \bibinfo{author}{\bibfnamefont{J.~W.} \bibnamefont{York{,}~Jr.}}
2203: (\bibinfo{year}{2002}), \bibinfo{note}{in preparation}.
2204:
2205: \bibitem[{\citenamefont{Baumgarte et~al.}(1996)\citenamefont{Baumgarte, Cook,
2206: Scheel, Shapiro, and Teukolsky}}]{Baumgarte-Cook-etal:1996}
2207: \bibinfo{author}{\bibfnamefont{T.~W.} \bibnamefont{Baumgarte}},
2208: \bibinfo{author}{\bibfnamefont{G.~B.} \bibnamefont{Cook}},
2209: \bibinfo{author}{\bibfnamefont{M.~A.} \bibnamefont{Scheel}},
2210: \bibinfo{author}{\bibfnamefont{S.~L.} \bibnamefont{Shapiro}},
2211: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{S.~A.}
2212: \bibnamefont{Teukolsky}}, \bibinfo{journal}{Phys. Rev. D}
2213: \textbf{\bibinfo{volume}{54}}, \bibinfo{pages}{4849} (\bibinfo{year}{1996}).
2214:
2215: \bibitem[{\citenamefont{Cook}(1994)}]{Cook:1994}
2216: \bibinfo{author}{\bibfnamefont{G.~B.} \bibnamefont{Cook}},
2217: \bibinfo{journal}{Phys. Rev. D} \textbf{\bibinfo{volume}{50}},
2218: \bibinfo{pages}{5025} (\bibinfo{year}{1994}).
2219:
2220: \bibitem[{\citenamefont{Brandt et~al.}(2000)\citenamefont{Brandt, Corell,
2221: G\'{o}mez, Huq, Laguna, Lehner, Marronetti, Matzner, Neilsen, Pullin
2222: et~al.}}]{Brandt-Corell-etal:2000}
2223: \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Brandt}},
2224: \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Corell}},
2225: \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{G\'{o}mez}},
2226: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Huq}},
2227: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Laguna}},
2228: \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Lehner}},
2229: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Marronetti}},
2230: \bibinfo{author}{\bibfnamefont{R.~A.} \bibnamefont{Matzner}},
2231: \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Neilsen}},
2232: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Pullin}},
2233: \bibnamefont{et~al.}, \bibinfo{journal}{Phys. Rev. Lett.}
2234: \textbf{\bibinfo{volume}{85}}, \bibinfo{pages}{5496} (\bibinfo{year}{2000}).
2235:
2236: \bibitem[{\citenamefont{Damour et~al.}(2002)\citenamefont{Damour, Gourgoulhon,
2237: and Grandcl{\'e}ment}}]{Damour-Gourgoulhon-Grandclement:grqc0204011}
2238: \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Damour}},
2239: \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Gourgoulhon}},
2240: \bibnamefont{and}
2241: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Grandcl{\'e}ment}},
2242: \bibinfo{journal}{gr-qc/0204011} (\bibinfo{year}{2002}).
2243:
2244: \bibitem[{\citenamefont{Cook}(2002)}]{Cook:2001}
2245: \bibinfo{author}{\bibfnamefont{G.~B.} \bibnamefont{Cook}},
2246: \bibinfo{journal}{Phys. Rev. D} \textbf{\bibinfo{volume}{65}},
2247: \bibinfo{pages}{084003} (\bibinfo{year}{2002}).
2248:
2249: \end{thebibliography}
2250:
2251:
2252: \end{document}
2253: