gr-qc0303099/eh.tex
1: %%%%%%%%%%%%%%%%%% file template.tex %%%%%%%%%%%%%%%%%%%%
2: %                                                       %
3: %    Copyright (c) Optical Society of America, 1992.    %
4: %                                                       %
5: %%%%%%%%%%%%%%%%%%% November 17, 1992 %%%%%%%%%%%%%%%%%%%
6: %
7: % THIS FILE IS A TEMPLATE TO PRODUCE AN ARTICLE SUBMISSION
8: % TO THE OSA JOURNALS, JOSA-A, JOSA-B, and APPLIED OPTICS.
9: %
10: % THIS TEMPLATE CONTAINS TYPESETTING COMMANDS WHICH BEGIN WITH A
11: % BACKSLASH.  THESE COMMANDS WILL BE READ BY LATEX, USING THE
12: % REVTEX 3.0 STANDARD MACROS.   PLEASE FILL IN THE REQUIRED DATA
13: % FOR THE MACROS, BUT DO NOT ALTER THE DEFINITIONS.
14: %
15: % EXAMPLE: IN \author{Authors' names} , PLEASE FILL IN THE
16: % AUTHORS' NAME(S).
17: %
18: % COMMENTS BEGIN WITH THE PERCENT (%) SYMBOL. AFTER A %, ANY
19: % DATA ON THE REST OF A LINE WILL NOT PRINT.
20: %
21: \documentstyle[twocolumn,prd,aps]{revtex}  % DON'T CHANGE
22: %
23: %
24: \newcommand{\MF}{{\large{\manual META}\-{\manual FONT}}}
25: \newcommand{\manual}{rm}        % Substitute rm (Roman) font.
26: \newcommand\bs{\char '134 }     % add backslash char to \tt font
27: %
28: %
29: \input epsf
30: \begin{document}                % INITIALIZE - DONT CHANGE
31: %
32: %
33: %
34: \title{Tracking Black Holes in Numerical Relativity}
35: \author{Scott A. Caveny, Matthew Anderson, Richard A. Matzner}
36: \address{Center for Relativity, The University of Texas at Austin, Austin, 
37: TX 78712-1081 }
38: % \author{}   % Use this and the next line only if there is a second
39: % \address{Another University, etc.}  % address. (Remove the left % marks)
40: %
41: \maketitle
42: \begin{abstract}                % DON'T CHANGE THIS LINE
43: This work addresses and solves the problem of generically tracking
44: black hole event horizons in computational simulation of black hole
45: interactions. Solutions
46: of the hyperbolic eikonal equation, solved on a curved spacetime
47: manifold containing black hole sources, are employed in
48: development of a robust tracking method capable of continuously
49: monitoring arbitrary changes of topology in the event horizon, as
50: well as arbitrary numbers of gravitational sources. The method
51: makes use of continuous families of level set viscosity solutions
52: of the eikonal equation with identification of the black hole
53: event horizon obtained by the signature feature of discontinuity
54: formation in the eikonal's solution. The method is employed in the
55: analysis of the event horizon for the asymmetric merger in a
56: binary black hole system. In this first such three dimensional analysis, we
57: establish both
58: qualitative and quantitative physics for the asymmetric collision;
59: including: 1. Bounds on the topology of the throat connecting the
60: holes following merger, 2. Time of merger, and 3. Continuous
61: accounting for the surface of section areas of the black hole
62: sources.
63: \end{abstract}
64: %
65: \section{Introduction}               % Introduction goes below.
66: The idea of a totally collapsed gravitational source, from which
67: nothing --- not even light --- can escape, is an old one, dating
68: back at least to the work of Laplace (and others) in the
69: eighteenth century \cite{he:}. Since those original
70: considerations, research has uncovered a wealth of understanding
71: regarding the physics of black holes. In general, the work has
72: followed two major routes,  with an ever - narrowing gap between
73: the approaches. Along one direction, black holes are studied as
74: mathematical solutions in a given theory of gravity including
75: Newtonian theory, post-Newtonian metric theories of gravity,
76: Einsteinian gravity, as well as semi-classical and (more recently)
77: various attempts at full quantum theories of gravity. The second
78: direction considers black holes in the astrophysical and
79: astronomical contexts. Both fields of research have seen
80: increasing activity over the years and have made startling discoveries
81: concerning the physics of black holes; including the gravitational
82: collapse theory of Oppenheimer  and Snyder \cite{opsny:}, proofs
83: of uniqueness and stability of black holes in Einstein's general
84: relativity \cite{carter:}, thermodynamic properties of black holes
85: including the three laws of black hole mechanics \cite{bch:} and
86: mechanisms for black hole radiation and evaporation \cite{wald2:},
87: discovery of critical phenomena in black hole formation
88: \cite{choptuik:}, experimental identification of both
89: astrophysical black hole sources themselves \cite{as1:} and their
90: event horizons \cite{e1:}, experimental signatures for super -
91: massive black holes in galactic centers \cite{ambh:}, and
92: experimental bounds on the distribution and spectrum of black hole
93: sources and collisions \cite{eb:}.
94: 
95: 
96: Numerical relativity, which addresses computational solution of
97: Einstein's equation \cite{lehner:}, is an active participant in
98: both the mathematical and the astrophysical programmes of research. 
99: With the advent of the Laser
100: Interferometric Gravitational Observatory and other similar efforts, 
101: considerable attention has focused on the
102: generic binary black hole coalescence (BBHC) problem, expected as
103: the strongest sources for gravitational waves \cite{bbhc1:}. As
104: discussed in \cite{lehner:}, this is a very difficult and
105: interesting problem that can only successfully be addressed in the
106: computational domain. Consequently, the binary black hole
107: coalescence problem has become an active subject of numerical
108: relativity. One particular problem in the computational domain is
109: the computational definition, detection and tracking of the black
110: hole event horizon itself \cite{eh1:}, \cite{eh2:}, \cite{eh3:},
111: \cite{eh4:}, \cite{eh5:}, \cite{is1:}.
112: 
113: The present work completely solves the problem of numerically
114: tracking black hole event horizons. The solution is complete in
115: the sense that a single method is presented such that any one
116: implementation of the method can generically detect arbitrary
117: numbers of black hole event horizons undergoing arbitrarily strong
118: gravitational interactions. For example, using a single computational
119: code of our method we analyse both single black holes and black holes
120: undergoing merger; and no special modifications of our code are 
121: required to handle these distinct dynamics. 
122: 
123: The present article, describing our generic method 
124: for tracking black hole event horizons, is divided as follows. In section II
125: the eikonal equation, the foundation of our method, is described in
126: sufficient detail to be employed in an event horizon tracker. In particular,
127: we focus on the signature behavior of a black hole event 
128: horizon in solutions of the eikonal equation. As usual in numerical work,
129: there are a variety of possible implementations of the approach. In section
130: III, several closely related systems of equations are presented and one 
131: particular system is singled out for consideration. The system chosen
132: makes use of
133: an explicit second order diffusion, or viscosity, term and we show in section 
134: III the relationship between solutions of the diffusive equations of motion
135: and the continuum eikonal equation of interest. Section IV details
136: extraction of the two dimensional sections of the event horizon for each 
137: time level of a numerical evolution. Such extraction is crucial for 
138: carrying out area, mass, and spin calculations. Section IV also shows 
139: the accuracy of our implementation by considering a parameter space survey 
140: of single Kerr black holes. Sections V - X describe in detail
141: the first three dimensional application of our (or any) method
142: to the numerical analysis of the asymmetric
143: binary black hole coalescence problem. In particular we search for
144: evidence of any nontrivial topology of the horizon immediately following 
145: merger.
146: In distinction to the prediction of \cite{wini2:} we find a topologically
147: spherical horizon to within the accuracy of our three dimensional mesh.
148: Further, area analysis of the candidate numerical event horizon is 
149: carried out in conjunction with analysis of the black hole 
150: apparent horizons, which reveals both a time of merger much earlier then 
151: estimated using apparent horizons and  mass energy estimates much 
152: larger those found using apparent horizon tracking methods. Our conclusions
153: are presented in section XI.
154: 
155:  
156: 
157: \section{The Eikonal}
158: To begin, since the Lagrangian $L = g_{ab}\dot{x}^{a}\dot{x}^{b}$
159: of null geodesic motion has only kinetic terms it is equal to the
160: associated Hamiltonian\footnote{We use early-latin indices to denote
161: spacetime components $a,b,c = 0,1,2,3$; mid-latin indices denote
162: spatial components $i,j,k=1,2,3$.}. Legendre transformation
163: \begin{equation}
164: H\left(\tau, x^{a},p_{b}\right) =\frac{d{x}^{c}}{d\tau} p_{c} -
165: L\left(\tau, x^{d},\frac{d{x}^{e}}{d\tau}\right)
166: \end{equation}
167: where
168: \begin{equation}
169: p_{a} \equiv \frac{\partial L}{\partial
170: \left(\frac{d{x}^{a}}{d\tau}\right)},
171: \end{equation}
172: sets $L = H$. The corresponding Hamiltonian equations are
173: \begin{equation}
174: \frac{d{x}^{a}}{d\tau} = \frac{\partial H}{\partial p_{a}} = 2
175: g^{ab}p_{b}
176: \end{equation}
177: \begin{equation}
178: \frac{d{p}_{b}}{d\tau} =  - \partial_{b} H = -
179: p_{c}p_{d}\partial_{b} g^{cd}.
180: \end{equation}
181: 
182: It is generic that the spacetime metric is independent of the
183: affine parameter $\tau$. There is thus a first integral associated
184: to geodesic motion. Making use of this property permits
185: elimination of the affine parameter in favor of coordinate time
186: $t$. To see this, it is convenient to adopt the ADM variables
187: \begin{equation}
188: g^{tt} = \frac{1}{\alpha^{2}},g^{ti} =
189: \frac{\beta^{i}}{\alpha^{2}},g^{ij} = \gamma^{ij} -
190: \frac{\beta^{i}\beta^{j}}{\alpha^{2}}.
191: \end{equation}
192: Here $\gamma^{ij}$ is the inverse spatial metric. 
193: Substituting the ADM variables directly into the Hamiltonian gives
194: \begin{equation}\label{6eq}
195: H=p_{a}g^{ab}p_{b} = -\frac{1}{\alpha^{2}}p_{t}^{2}
196: +\frac{2}{\alpha^{2}}\beta^{i}p_{i} + p_{i}\left(\gamma^{ij} -
197: \frac{1}{\alpha^{2}} \beta^{i}\beta^{j} \right)p_{j}.
198: \end{equation}
199: Since the Hamiltonian is not explicitly $\tau$ dependent  there is
200: a constant of the motion
201: \begin{equation}\label{geodesics}
202: -\frac{1}{\alpha^{2}}p_{t}^{2} +\frac{2}{\alpha^{2}}\beta^{i}p_{i}
203: + p_{i}\left(\gamma^{ij} - \frac{1}{\alpha^{2}} \beta^{i}\beta_{j}
204: \right)p_{j} = \omega^{2}.
205: \end{equation}
206: For $\omega^{2} < 0 $, $\omega^{2} =0 $, or $\omega^{2} > 0 $ the
207: motion is said to be timelike, null, or spacelike. Without loss of
208: generality,  assuming null geodesic motion, solution of
209: (\ref{geodesics}) by ordinary algebra yields:
210: \begin{equation}\label{p}
211: p_{t} = \beta^{i} p_{i} \pm \alpha \sqrt{p_{i}\gamma^{ij}p_{j}}.
212: \end{equation}
213: With this result, the Hamiltonian can be explicitly factored:
214: \begin{equation}
215: H = H_{+}H_{-} = 0
216: \end{equation}
217: where
218: \begin{equation}\label{root}
219: H_{\pm} = p_{t} - \beta^{i}p_{i} \pm \alpha
220: \sqrt{p_{i}\gamma^{ij}p_{j}} = 0.
221: \end{equation}
222: In the case of either root Hamilton's canonical equations become
223: \begin{equation}\label{teq}
224: \frac{dt}{d\tau} = H_{\mp}\frac{\partial H_{\pm}}{\partial p_{t}}
225: = H_{\mp}
226: \end{equation}
227: \begin{equation}\label{2eq}
228: \frac{dx^{i}}{d\tau} = H_{\mp}\frac{\partial H_{\pm}}{\partial
229: p_{i}} = H_{\mp}\left(-\beta^{i} \pm \alpha
230: \frac{\gamma^{ij}p_{j}}{\sqrt{p_{k}\gamma^{kl}p_{l}}}\right)
231: \end{equation}
232: \begin{equation}
233: \frac{dp_{t}}{d\tau} = - H_{\mp}\frac{\partial H_{\pm}}{\partial
234: t}
235: \end{equation}
236: \begin{equation}\label{4eq}
237: \frac{dp_{i}}{d\tau} = - H_{\mp}\frac{\partial H_{\pm}}{\partial
238: x^{i}}.
239: \end{equation}
240: According to equation (\ref{teq}), $\tau$ can be eliminated in
241: favor of $t$. The system of equations becomes simply the equations
242: (\ref{2eq})
243: and (\ref{4eq})  with $t$ written in place of  $\tau$ and the
244: factor $H_{\mp}$ cancelled everywhere on the right hand side.
245: Explicitly,
246: \begin{equation}\label{null1}
247: \frac{dx^{i}}{dt} =-\beta^{i} \pm \alpha
248: \frac{\gamma^{ij}p_{j}}{\sqrt{p_{k}\gamma^{kl}p_{l}}}
249: \end{equation}
250: and
251: \begin{equation}\label{null2}
252: \frac{dp_{i}}{dt} = - \partial_{i}\left(- \beta^{j}p_{j} \pm
253: \alpha \sqrt{p_{j}\gamma^{jk}p_{j}}\right).
254: \end{equation}
255: The eikonal, corresponding to the Hamiltonian of equation (\ref{6eq}),
256: \begin{equation}
257: \partial_{a}Sg^{ab}\partial_{b}S = 0
258: \end{equation}
259: can be factored similarly. To do so it is a simple matter of
260: making the replacements $p_{t} \rightarrow
261: \partial_{t}S $, $p_{i} \rightarrow \partial_{i}S $ in (\ref{p})
262: to find the following symmetric hyberbolic partial differential
263: equation
264: \begin{equation}\label{PDE}
265: \partial_{t}S = \beta^{i} \partial_{i} S \pm \alpha\sqrt{\partial_{i}S
266: \gamma^{ij} \partial_{j}S} \equiv \bar{H}.
267: \end{equation}
268: Note that a bar is introduced here to distinguish the Hamiltonian
269: used here from the Hamiltonian used in (\ref{6eq}). 
270: This equation is also used in the method of \cite{eh1:}, although
271: in that work the equation is further reduced to consider the case of 
272: a single null surface. 
273: 
274: 
275: The right hand side of this result is homogeneous of degree one in
276: $\partial_{i}S$.  The characteristic curves along which the level
277: sets $\Gamma$ of $S$ are propagated, are then
278: \begin{equation}\label{x2}
279: \dot{x}^{i} = -\beta^{i} \pm \alpha \frac{\partial^{i}S} {
280: \sqrt{\partial_{i}S \gamma^{ij} \partial_{j}S}} \equiv
281: \frac{\partial \bar{H}}{\partial\left(\partial_{i} S\right)}
282: \end{equation}
283: \begin{equation}\label{p2}
284: \partial_{i}\dot{S} = \partial_{i} \bar{H}\left(t,x^{j},\partial_{j}S\right),
285: \end{equation}
286: which are the null geodesics of equations (\ref{null1}) and (\ref{null2}). 
287: Immediately, the integral curves of
288: the gradients of $S$ and $\Gamma$ are also the null geodesics:
289: \begin{equation}
290: \frac{dx^{i}\left(\lambda\right)}{d\lambda} = g^{i a} p_{a}  = g^{i a} 
291: \partial_{a} S = \partial^{i}
292: S\left(\lambda, x^{j}\left(\lambda\right)\right).
293: \end{equation}
294: 
295: Hereafter, the bar on $\bar{H}$
296: is dropped with the understanding that the Hamiltonian considered is 
297: that of equation (\ref{PDE}).
298: This result establishes that the eikonal is technically a Riemann
299: invariant of the null geodesics, a fact that proves useful
300: in establishing the signature of a black hole event horizon
301: in solutions of the eikonal equation. More specifically, since 
302: the eikonal is the canonical generator of null geodesics, it
303: can be employed in
304: analysis of black hole event horizons, which are 
305: by definition generated by null
306: geodesics having no future end points. To proceed in this
307: manner the equation (\ref{PDE}) is employed in an initial value
308: problem and then surveyed for signature features of black hole
309: event horizons.
310: 
311: A black hole event horizon is generated by 
312: a congruence of outgoing --- but future asymptotically nonexpanding
313: --- null geodesics. The scope of the surveys of the eikonal equation
314: that are  required to identify a black hole 
315: event horizon is then restricted to 
316: the space of all outgoing null surfaces. These surveys are
317: greatly reduced by the fact that solutions of the eikonal are
318: categorized by topologically equivalent solutions. To see this,
319: note that
320: \begin{equation}
321: \partial_{i}\psi\left(S\right) = \frac{\partial \psi}{\partial S}
322: \partial_{i} S = \lambda\left(S\right) \partial_{i}S,
323: \end{equation}
324: and
325: \begin{equation}
326: \partial_{t}\psi\left(S\right) = \frac{\partial \psi}{\partial S}
327: \partial_{t} S = \lambda\left(S\right) \partial_{t}S.
328: \end{equation}
329: By homogeneity of the right hand side of (\ref{PDE}), if $S$ is a
330: solution then $\psi\left(S\right)$ is also solution. Thus smoothly
331: related initial data $S_{0} \rightarrow S_{0}' =
332: \psi\left(S_{0}\right)$ have smoothly related solutions. This
333: feature alleviates the need for surveying over smoothly related
334: initial data.
335: 
336: A further reduction of the scope of solution surveys is provided
337: by the equivalence of ingoing and outgoing solutions under time
338: reversal. Propagation of data for $S$ describing an ingoing or outgoing 
339: null surface is
340: accomplished by specification of: 1. A definition of the direction
341: of time, 2. A choice of $\alpha$ and $\beta^{i}$, and 3. A choice
342: of the root. With these choices specified, data is then uniquely
343: partitioned into an ingoing type and an outgoing type with the
344: distinction being the gradient of $S$.
345: 
346: With the dynamics of any outgoing null surface (including the
347: event horizon) specified by (\ref{PDE}), and the scope of solution
348: surveys categorized into topological classes, the task remains of
349: identifying the event horizon within this restricted space of 
350: outgoing null surfaces. To do so, the results and approach
351: of \cite{eh1:} are adopted here 
352: and
353: modified to include the eikonal equation as discussed above. The
354: result of this approach is a signature feature of black hole event
355: horizons in the eikonal equation.
356: 
357: An event horizon of
358: a black hole is by definition a critical outgoing surface when
359: tracked into the future. Let $\mathcal{P}$ be a
360: point interior to the horizon and $\mathcal{Q}$ be a point
361: exterior to the horizon such that $\mathcal{P}$ and $\mathcal{Q}$
362: lie on characteristic curves of the eikonal $\gamma_{{\mathcal{P}}}$ and
363: $\gamma_{{\mathcal{Q}}}$. At arbitrarily early times let
364: $\gamma_{{\mathcal{P}}}$ and $\gamma_{{\mathcal{Q}}}$ pass
365: arbitrarily closely to a point $\mathcal{H}$ that lies on the
366: horizon $\Gamma$. Since $S$ is a Riemann invariant, at arbitrarily
367: early times the jump of the eikonal at $\mathcal{H}$ becomes
368: $\left[\left[S\left(\mathcal{H}\right)\right]\right] \equiv
369: S_{-}\left(\mathcal{H}\right) - S_{+}\left(\mathcal{H}\right)
370: \approx S\left(\mathcal{P}\right) - S\left(\mathcal{Q}\right) \neq
371: 0$. In the computational domain, where the resolution is finite,
372: this discontinuity will appear generically in finite time. As
373: such, an approximation of the  event horizon will appear
374: numerically as the formation of a jump discontinuity in the
375: eikonal for outgoing data that is propagated into the past.  This
376: is the numerical signature of black hole event horizons in the
377: eikonal's solutions.
378: 
379: \section{Numerical Methods}
380: Analysis of the continuum properties of the eikonal equation as a
381: Hamilton Jacobi equation identifies
382: three closely related approaches to tracking black hole event
383: horizons:
384: \begin{itemize}
385: \item System I: The null geodesic equations
386: \begin{equation}
387: S\left(x^{j},t\right) = S_{0}\left(x^{k}\right) - \int dt
388: H\left(t, x^{i}, p_{i}\right)
389: \end{equation}
390: where the integral is evaluated along the solutions of
391: \begin{eqnarray}
392: \dot{x}^{i} &=& \beta^{i} \pm \alpha
393: \frac{p^{i}}{\sqrt{p_{i}\gamma^{ij}p_{j}}},\\
394: \nonumber \dot{p}_{i} &=& -\partial_{i}H. \\ \nonumber
395: \end{eqnarray}
396: 
397: \item System II: The eikonal
398: \begin{equation}\label{eikonal}
399: \partial_{t} S = -H\left(t, x^{j}, \partial_{j}S\right)
400: \end{equation}
401: 
402: \item System III: The flux conservative form
403: \begin{equation}
404: S\left(x^{i},t\right) = \oint dx^{i}  p_{i}\left(x^{j},t\right)
405: \end{equation}
406: \begin{equation}\label{flux}
407: \partial_{t} p_{i}\left(x^{i},t\right)
408: + \partial_{i}H\left(t, x^{j}, p_{j}\right) = 0.
409: \end{equation}
410: \end{itemize}
411: In each of the above three systems of equations the Hamiltonian is
412: given by (\ref{PDE}).
413: 
414: In each of the above cases, the structure of the dynamical
415: equations provide certain advantages. For example, in each case the
416: symplectic structure can be employed to identify 
417: numerical loss of accuracy in a similar manner to some modern
418: numerical schemes used in Hamiltonian dynamics. 
419: Further, as a flux conservative system,
420: high resolution methods from computational fluid dynamics can be
421: applied directly to the third system. In following sections of this
422: article, the
423: second system is considered since this system of equations 
424: yields an expedient implementation that is sufficiently accurate
425: for our purposes.
426: 
427: Singular behavior on the eikonal is not specific to only the event
428: horizon and instead, as described in detail by Arnold and Newman
429: \cite{newman1:}, \cite{newman2:}, \cite{newman3:},
430: the eikonal is known to generically break down on caustic and
431: other sets. Special numerical methods are then required to handle
432: the generic singular behavior of the eikonal; we make use of an explicit viscosity term.  In the
433: continuum, addition of our form of numerical viscosity at the level of the
434: finite difference approximation corresponds to replacing the
435: evolution of
436: \begin{equation}
437: \partial_{t} S = - H\left(t,x^{i},\partial_{j}S\right)
438: \end{equation}
439: with evolution of the equation
440: \begin{equation}\label{WKB}
441: \partial_{t} \psi =
442: \epsilon^{2} \nabla^{2} \psi -
443: H\left(t,x^{i},\partial_{j}\psi\right)
444: \end{equation}
445: where $\epsilon$ is a small quantity which we call the viscosity
446: and $\nabla^{2}$ denotes any
447: second order, linear derivative operator. There is a well defined
448: sense in which the solutions $S$ relate to the solutions $\psi$;
449: it is simply given (when the solutions $S$ exist) by the WKB
450: transformation
451: \begin{equation}
452: \psi\left(x^{i},t\right) = \sum_{n}
453: a_{n}\left(x^{i},t\right)\epsilon^{n}
454: \exp\left(S/\epsilon\right)\equiv A \exp(S/\epsilon).
455: \end{equation}
456: To see the explicit relationship between  the solutions $\psi$ and
457: the solutions $S$ note that
458: \begin{equation}
459: \partial_{t} \psi =
460: \frac{\psi}{\epsilon}\left(\partial_{t} S + \epsilon
461: \partial_{t} \log A\right),
462: \end{equation}
463: \begin{equation}
464: \partial_{i} \psi =
465: \frac{\psi}{\epsilon}\left(\partial_{i} S + \epsilon
466: \partial_{i} \log A\right),
467: \end{equation}
468: \begin{equation}
469: \nabla^{2} \psi = \frac{\psi}{\epsilon^{2}} \left( \nabla S +
470: \epsilon \nabla \log A \right)^{2} +
471: \frac{\psi}{\epsilon}\left(\nabla^{2}S  + \epsilon \nabla^{2} \log
472: A\right).
473: \end{equation}
474: Assuming that the Hamiltonian is homogeneous of degree one in
475: momentum, and making use of perturbation theory:
476: \begin{equation}
477: H\left(x^{i},\nabla_{j}\psi\right) = \frac{\psi}{\epsilon}\left(
478: H\left(x^{i},\nabla_{j}S\right) + \epsilon
479: H_{1}\left(t,x^{i},\partial_{j}\log A\right)\right).
480: \end{equation}
481: Substituting these results into (\ref{WKB}) and cancelling an
482: overall factor of $\psi/\epsilon$ gives at lowest, and first
483: order:
484: \begin{equation}
485: \partial_{t}S  =
486: - H\left(t,x^{i},\partial_{j}S\right).
487: \end{equation}
488: \begin{eqnarray}
489: \epsilon \partial_{t} \log A  &=& - \epsilon
490: H_{1}\left(t,x^{i},\partial_{j}\log A\right)\\ \nonumber &+&
491: \epsilon \left(\nabla S + \epsilon \nabla \log A \right)^{2} +
492: \ldots \\ \nonumber
493: \end{eqnarray}
494: Here $H_{1}$ is the first order linear Hamiltonian obtained from
495: perturbation theory. At zeroth order, $S$ then satisfies the
496: eikonal equation, while the first order correction is the linear
497: result of first order perturbation theory and expresses the
498: evolution of $A$. Higher order results can be found similarly.
499: Again, these results hold only where solutions of $S$ exist; that
500: is, away from discontinuities or other solution singularities in
501: $S$ and its derivatives.
502: 
503: The above analysis is a slight modification of usual WKB expansion
504: and expresses the solutions $\psi$ in terms of the solutions $S$.
505: Such solutions can be inverted using the conventional method of
506: series inversion. To invert the WKB expression, and express the
507: solutions of the eikonal in terms of the solution of the parabolic
508: equation, assume first that the solution of the parabolic equation 
509: (\ref{WKB})
510: is given as $\psi$ (numerically or otherwise). Writing
511: \begin{equation}\label{spsi}
512: S = \log \left(\frac{\psi}{\bar{A}}\right),
513: \end{equation}
514: the procedure is then analogous to that of the WKB expansion:
515: $\bar{A}$ is constructed as an asymptotic power series with
516: coefficients depending on $\psi$ alone. The result of such an
517: analysis is simply:
518: \begin{equation}
519: S\left(x,t\right) = \log \psi\left(x,t\right)  + \bar{\epsilon}
520: \int dt' \frac{\nabla^{2}\psi\left(\bar{x},t\right)}{\psi}
521: \end{equation}
522: where the integral is evaluated along
523: \begin{equation}
524: \frac{d\bar{x}^{i}}{dt} = \frac{\partial
525: H\left(t,\bar{x}^{j},\partial_{\bar{x}^{k}}\log\psi\right)}{\partial
526: p_{i}}
527: \end{equation}
528: and $\psi$ is provided independently by (numerical) solution of
529: \begin{equation}
530: \partial_{t} \psi = \bar{\epsilon} \nabla^{2}\psi -
531: H\left(t,x^{i},\partial_{j}\psi\right).
532: \end{equation}
533: 
534: One advantage of an explicit second order viscosity term is a
535: simple procedure for reducing the error of viscosity solutions by one
536: order of the viscosity $\epsilon$.  The limit $\epsilon \rightarrow 0$ corresponds to a continuous
537: family of zeroth order solutions $\phi_{\epsilon}$; 
538: where 
539: \begin{equation}
540: \phi_{\epsilon} = S + \eta_{\epsilon}
541: \end{equation}
542: and $\eta_{\epsilon} = {\mathcal{O}}\left(\epsilon\right)$. Note that
543: each solution $\phi_{\epsilon}$ is obtained from an analysis similar
544: to that following equation (\ref{spsi}), although obtained
545: by neglecting the logarithm. 
546: 
547: Given two viscosity solutions $\phi_{\epsilon}$ and $\phi_{2
548: \epsilon}$ it is then possible to construct a third improved
549: viscosity solution $\phi_{I}$ that is accurate to
550: ${\mathcal{O}}\left(\epsilon^{2}\right)$. To see this, consider the
551: combination
552: \begin{equation}
553: \phi_{I} \equiv \phi_{\epsilon}  + \left(\phi_{\epsilon} - \phi_{2
554: \epsilon}\right).
555: \end{equation}
556: $\phi_{I}$ is accurate to
557: ${\mathcal{O}}\left(\epsilon^{2}\right)$ since
558: \begin{equation}
559: \phi_{\epsilon} - \phi_{2 \epsilon} = \eta_{\epsilon} - \eta_{2
560: \epsilon} \approx  \eta_{\epsilon} - 2 \eta_{\epsilon} = -
561: \eta_{\epsilon}.
562: \end{equation}
563: 
564: To make use of these improved viscosity solutions  let $h$ denote
565: the resolution of the numerical mesh. Any second order finite
566: difference approximation will have then
567: \begin{equation}
568: \phi = \hat{\phi} + {\mathcal{O}}\left(h^{2}\right),
569: \end{equation}
570: where $\phi$ is the continuum solution and $\hat{\phi}$ denotes
571: its finite difference approximation.[We will use a hat  throughout
572: the paper to denote the discrete approximation to a continuum object.] Similarly,
573: \begin{equation}
574: S = \hat{S} + {\mathcal{O}}\left(h^{2}\right).
575: \end{equation}
576: Using
577: \begin{equation}
578: S =\phi + {\mathcal{O}}\left(\epsilon\right) =  \phi_{I} +
579: {\mathcal{O}}\left(\epsilon^{2}\right)
580: \end{equation}
581: gives
582: \begin{equation}
583: \hat{S}= \hat{\phi} + { \mathcal{O}}\left(h^{2}\right) +
584: {\mathcal{O}}\left(\epsilon\right) = \hat{\phi_{I}} + {
585: \mathcal{O}}\left(h^{2}\right) +
586: {\mathcal{O}}\left(\epsilon^{2}\right).
587: \end{equation}
588: 
589: \section{Level Set Extraction}
590: Of crucial interest in the binary black hole coalescence problem
591: are the areas of sections of the black hole event horizon.
592: To find these sections, at any given time level  
593: any of the level sets, say $S =
594: 0$, can be extracted to obtain the surface
595: $\Gamma$, a two dimensional section of the corresponding null surface. 
596: This problem of extraction is an inverse problem, since
597: it requires that points $\left(x,y,z\right)$ are found such that
598: $S\left(x,y,z\right) = 0$. To accomplish this inversion, we find that
599: an ordinary
600: bisection method is sufficient for use with an ordinary second order
601: interpolation scheme. In this method it is assumed that the
602: surface $\Gamma$ can be expressed in spherical coordinates
603: $\left(\theta, \phi, u\left(\theta, \phi\right)\right)$, where $r
604: = u\left(\theta, \phi\right)$ is the surface function for a given
605: center $c^{i}$ contained within the surface $\Gamma$. Given a
606: choice for the center, the radial function for
607: $S\left(\Gamma\right) = S_{o} = 0$ can then be approximated via the interpolation
608: and bisection.
609: 
610: To establish the accuracy of our implementation, including
611: the routines that accomplish 
612:  extraction of  the level sets and the accuracy of 
613: the viscosity term, we consider stationary, spinning black holes. 
614: This case is
615: completely described by the Kerr - Newmann family of axisymmetric solutions of Einstein's equation. It is convenient to make
616: use of the Kerr-Schild form for the metric \cite{deirdre:} since 
617: this form is used in our binary black hole evolution code as well
618: as in our solution of the initial value problem for setting
619: initial data for the evolution. Specifically,
620: the metric in the Kerr-Schild form is
621: \begin{equation}
622: g^{ab} = \eta^{ab} - 2 H l^{a}l^{b}.
623: \end{equation}
624: Here $\eta_{ab} = \mathrm{diag}\left(-1,1,1,1\right)$ is
625: Minkowski's metric, $H$ is a space time scalar, and $l^{a}$ is an
626: ingoing null vector with respect to both the Minkowski and full
627: metric. The Kerr solution is the two parameter family of solutions
628: such that
629: \begin{equation}
630: H = \frac{M r^{3}}{r^{2} + a^{2} z^{2}}
631: \end{equation}
632: and
633: \begin{equation}
634: l^{t} = -1,
635: \end{equation}
636: \begin{equation}
637: l^{x} = \frac{r x + a y}{r^{2} + a^{2}},
638: \end{equation}
639: \begin{equation}
640: l^{y} = \frac{r y - a x}{r^{2} + a^{2}},
641: \end{equation}
642: \begin{equation}
643: l^{z} = \frac{z}{r},
644: \end{equation}
645: \begin{equation}
646: r^2 = \frac{1}{2}\left(\rho^{2} - a^{2}\right) +
647: \sqrt{\frac{1}{4}\left(\rho^{2} - a^{2}\right) + a^{2} z^{2}}
648: \end{equation}
649: where
650: \begin{equation}
651: \rho^2 = x^{2} + y^{2} + z^{2}.
652: \end{equation}
653: Finally, the event horizon for the Kerr black hole is located
654: on the ellipsoid $r = r_{+} = r\left(x,y,z\right)$ where
655: \begin{equation}
656: r_{+} = M + \sqrt{M - a}.
657: \end{equation}
658: \begin{figure} % Imported eps example.
659: \epsfxsize=8cm
660: \centerline{\epsfbox{verrora.ps}}
661: \vspace{0.5cm}
662: \caption{Percent error in area of $M = 2$, $a = 0$ event
663: horizons in survey over viscosity parameter: $\epsilon = $
664: $h^2$, ${h^2}/2$, ${h^2}/4$, ${h^2}/8$. Here increasing $t$ corresponds
665: to propagation into the past.} \label{verrora:1}
666: \end{figure}
667: 
668: In figures (\ref{verrora:1}) -(\ref {vsrmse:1}) we show the evolution 
669: backward in time of the eikonal equation (followed by extraction of the $S=0$ 
670: surface). Errors can arise both in the evolution and in the extraction of the 
671: surface. Figure (\ref{verrora:1}) shows the percent errors in the extracted 
672: areas for a nonspinning black holes with mass $M = 2$, $a = 0$ in a survey over 
673: the viscosity parameter. Figure (\ref{vrmse:1}) shows a similar study but for 
674: the L2 norm of the truncation error in the function $r_{+}$. Note that 
675: the function $r_{+}$ is defined for every point of the horizon in the
676: continuum $r_{+} = r_{+}\left(x,y,z\right)$. In the computational domain
677: $r_{+}$ then takes a discrete form 
678: $\hat{r}_{+} = \hat{r}_{+}\left(i,j,k\right)$, where the integers $i,j,k$ span
679: the numerical mesh. The truncation error $e_{r_{+}}$ is then
680: \begin{equation}
681: e_{r_{+}}\left(i,j,k\right) = r_{+}\left(x,y,z\right) - \hat{r}_{+}\left(i,j,k\right)
682: \end{equation}
683: where is it understood that both $r_{+}$ and $\hat{r}_{+}$ are evaluated
684: at the same point. Bith Figures(\ref{verrora:1}) and (\ref{vrmse:1})show that 
685: viscosity parameter $h^2/8$ adequately captures the horizon location. While not 
686: perfect, it will suffice for the short term horizon tracking reported
687: here.
688: \begin{figure} % Imported eps example.
689: \epsfxsize=8cm
690: \centerline{\epsfbox{vrmse.ps}}
691: \vspace{0.5cm}
692: \caption{L2 norm in truncation error of $r_{+}$ for $M = 2$, $a = 0$ event
693: horizons in survey over viscosity parameter: $\epsilon = $
694: $h^2$, ${h^2}/2$, ${h^2}/4$, ${h^2}/8$. Here increasing $t$ corresponds
695: to propagation into the past.} \label{vrmse:1}
696: \end{figure}
697: However these results also suggest that in vanishing viscosity the percentage error in
698: the calculated area is reduced toward a bias. This bias is partly
699: associated to the finite resolution of the computational mesh and
700: partly to accuracy of the extraction routine. These figures were
701: generated using a three dimensional computational domain of
702: $N^{3}$ points with $N = 121$. The outer boundaries are located at
703: $\left[-15M, +15M\right]$ in the $x,y,z$ directions. The
704: resolution of the finite difference mesh for these results is then
705: $h = M / 4$. Also, a Courant -  Friedrichs -  Lewy factor of
706: $\lambda = 1/4$ with an iterated Crank Nicholson scheme \cite{icn:}
707: was used
708: as the finite difference approximation of the evolution of the eikonal
709: equation (\ref{PDE}).
710: Neumann boundary conditions $\partial_{i} S = 0$ on the outer boundary 
711: are found
712: to be generically
713: sufficient conditions for stability of the method. The philosophy
714: here is that the primary interest is deep within the bulk of the
715: computational domain where the event horizon of the black hole is
716: located. The outer boundary is then treated only to the degree that the
717: evolution of the interior region remains stable. Further, the
718: interior of the black hole is excised from the computational
719: domain in a sphere of radius $r = r_{0} + 2 dx$, where $r_{0}$
720: denotes the radius of the Kerr - Newman ring curvature
721: singularity. Finally, the discrete surface $\hat{\Gamma}$ was
722: constructed using $m^{2}$ points with $m = 100$. At such a
723: resolution of the extracted surface any errors in the area must be
724: attributed to all of: the viscosity parameter, the extraction routine, and
725: the resolution of the underlying three dimensional grid.
726: 
727: \begin{figure} % Imported eps example.
728: \epsfxsize=8cm
729: \centerline{\epsfbox{vserrora.ps}}
730: \vspace{0.5cm}
731: \caption{Percent error in area for three dimensional
732: level set solutions $a/M =$, $0$, $1/4$, $1/2$, $3/4$. Here 
733: increasing $t$ corresponds
734: to propagation into the past.} \label{vserrora:1}
735: \end{figure}
736: 
737: Figures (\ref{vserrora:1}) and (\ref{vsrmse:1}) show the percent
738:  errors in the extracted areas
739: as well as the L2 Norm of the truncation error in the function
740: $r_{+}$ in a survey over the angular momentum parameter $a$.
741: Here the viscosity is $\epsilon = h^{2}$. 
742: These figures both show the
743: evolution from a single null sphere that is completely
744: exterior to the horizon and propagated into the past. 
745: Since the event horizon of a spinning black hole 
746: is elliptical in its geometry, the spherical data we 
747: have chosen for $t_{\downarrow} = 0$ corresponds 
748: to a percent error in the area and $r_{+}$ that varies 
749: with the spin parameter $a/M$ at 
750: $t_{\downarrow} = 0$, explaining why the curves do not 
751: intersect at $t_{\downarrow} = 0$. (Where appropriate we append 
752: a $\downarrow$ to $t$, thus: $t_{\downarrow}$, 
753: indicating evolution into the past;
754: we also sometimes use $t_{\uparrow}$ to emphasize that we mean 
755: the forward evolving, usual, time $t$. Thus 
756: $t_{\downarrow} = 0$ corresponds to the late time at which we begin to 
757: integrate into the past.) 
758: 
759: However, the errors should converge to a constant, which
760: is evident in each of the curves with $a<0.75$ in 
761: figures (\ref{vserrora:1}) and (\ref{vsrmse:1}). For $a=0.75$ the curve
762: does not converge, and we expect that different
763: choice of initial data {\it will} exhibit convergence. 
764: According to these results, the viscosity level set method does
765: indeed accurately and robustly detect the distorted outermost
766: event horizons of spinning black holes at least when $a<0.75$. 
767: Note that we study both
768: the accuracy of $r_{+}$ and the accuracy of the calculated
769: areas since the calculations seperately and together establish
770: the accuracy of our area calculation and of our detection of the 
771: event horizon. 
772: 
773: \begin{figure} % Imported eps example.
774: \epsfxsize=8cm
775: \centerline{\epsfbox{vsrmse.ps}}
776: \vspace{0.5cm}
777: \caption{L2 norm of truncation error in $r_{+}$ for three dimensional
778: level set solutions $a/M =$ $0$, $1/4$, $1/2$, $3/4$. Here increasing $t$ corresponds
779: to propagation into the past.} \label{vsrmse:1}
780: \end{figure}
781: 
782: \section{Asymmetric Binary Black Hole Coalescence}
783: Analysis of the event horizon for the binary black hole
784: coalescence problem in  the case of head on collision has been
785: considered in detail in \cite{eh4:}. The problem of the event
786: horizon for asymmetric, that is off axis, collision has only been
787: considered analytically \cite{wini2:} and prior to this work no
788: results for numerically generated sources have been analyzed.
789: Numerical evolution and analysis of an asymmetric  binary
790: black hole system was studied in \cite{exc1:}, but
791: at that time the question of the event horizon was not considered.
792: 
793: In this section the results of the previous sections are applied
794: to the first completely numerical analysis of the event horizon
795: for the case of asymmetric collision.
796: 
797: To begin, consider two black holes of mass $M = 1$ with aligned
798: spins in the positive $z$ direction of $a/M = 1/2$. The 
799: computational domain is a grid of
800: $N^{3}$ points with $N = 121$. The outer boundary is located at
801: $\pm 15 M$ and the holes are initially positioned at 
802: $\left(x,y,z\right) = \left(+6,+2,0\right)$, and 
803: $\left(x,y,z\right) = \left(-6,-2,0\right)$. This computational domain
804: is identical to the mesh used in  the previous section to analyze
805: the percent error in area calculations  of surfaces extracted from
806: the level set method. The percent error in the
807: calculation of the area of sections of the horizon should
808: have a magnitude of about $4-5\%$. Further, as an
809: order of magnitude estimate, in a flat spatial geometry; e.g., in
810: a Newtonian spacetime, the initial separation $s$ of the black
811: hole centers would be $s = \sqrt{12^2 + 4^2} \approx 12.64$. This
812: would seem to be an ample initial separation to guarantee that the
813: initial data corresponds to two distinct black holes. However, in
814: the case that the holes are nonspinning, each will have a
815: spherical event horizon of radius $r = 2M$. Assuming only marginal
816: distortion of the nonspinning event horizons due to spin effects,
817: which is an approximation that is justified by the properties of spin
818: $a/M= 1/2$ black holes, the nearest separation
819: between the two sections of the black hole event horizon
820: is then
821: approximately $s_{min} = 12.64 M  - 4 M  \approx 8.64M $. Again,
822: this approximation assumes a flat underlying geometry; and so can
823: be considered only as an order of magnitude estimate. These initial
824: data then appear to correspond to a separation of approximately two
825: nonspinning black hole diameters between the surfaces of section
826: of the black holes. 
827: 
828: The holes are boosted along the $x$
829: direction with speeds of $\pm c / 2$. This boost lengthens the
830: nearest separation $s_{min}$ of the holes due to Lorentz
831: contraction of the horizons.[In this coordinate system the horizons
832: undergo contraction in the direction of motion. For a single hole, 
833: the area of the horizon does not change under this boost.] The nearest 
834: coordinate separation $s_{min}$
835: between the holes is then expected to lie in the range $8.64 <
836: s_{min} < 12.64$.
837: 
838: 
839: The numerical evolution of this collision process is carried out
840: for approximately $10M$ of run time with a Courant factor of
841: $\lambda = dt / dx = 1/4$. The code is the Texas black hole 
842: evolution code, a derivative of the Agave code \cite{exc1:}. 
843: Apparent horizon finders
844: \cite{deirdre:} locate two distinct apparent horizons of area $A
845: \approx 50 M $ for the initial data and continue to do so until $t
846: = 8M $, when only a single apparent horizon of area $A \approx
847: 200M$ can be located. This single apparent horizon persists until
848: approximately $10M$, beyond which instability effects, stemming
849: from the outer boundary and the excision boundary, swamp the
850: solution.
851: 
852: \section{Data For the Eikonal}
853: While the run length of this asymmetric collision data is
854: relatively small in units of the black hole masses, the problem of
855: detecting the associated black hole event horizon, or horizons, is
856: not a small computational problem. For example, to analyze this data requires analysis of the
857: lapse $\alpha$, the shift vector $\beta^{i}$, and the three metric
858: $\gamma^{ij}$. By symmetry of the metric, $\gamma^{ij} =
859: \gamma^{ji}$, there are only 6 independent components. Tracking
860: the event horizon of the associated data then requires analysis of
861: 10 grid functions, where each grid function consists of
862: ${\mathcal{O}}\left(Nt\times N^{3}\right)\times 8 B  =
863: {\mathcal{O}}\left(160\times121^{3}\right) \times 8 B \approx 2.26
864: GB$ of data. That is, tracking the associated event horizon
865: requires analysis of approximately $20 GB$ of data.
866: 
867: A more serious difficulty associated to this data set is the
868: relaxation time, $t\approx 4 M $, that is typically required for outgoing
869: data to converge onto the event horizon when followed into the
870: past. Assuming that the collision time is (as suggested by the
871: apparent horizon solvers) near $t = 8M$, perturbation theory
872: implies that the resulting horizon should undergo quasi normal
873: ringing for another $t \approx 20M$ from that time. That is, at
874: the time level $t = 10M$, where the event horizon tracker will
875: begin tracking into the past, it is expected that the event
876: horizon remains highly distorted and far from its stationary
877: regime. The problem is to determine good data for the eikonal at
878: the time level $t = 10M$, which can then be propagated into the
879: past. In contrast, in analysis of the event horizon for the case
880: of head on collision, researchers made use of approximately $100M$
881: of data \cite{eh4:}, and assumed that the final state at $t = 100M$
882: was a
883: stationary or quasi stationary black hole. In such a circumstance
884: using spherically symmetric data at $t = 100M$ for an event
885: horizon tracking method should prove sufficient.  The difficulty
886: with a data set only of length $t = 10M$ is that using spherically
887: symmetric data for the eikonal at the time level $t = 10M$ does
888: not accurately approximate sections of the event horizon at that
889: time.
890: 
891: 
892: To accommodate data of this $10M$ length and to generate
893: sufficient data for the eikonal at the time level $t = 10M$,  a
894: method for finding a candidate section of the event horizon at $t
895: = 10M$ is proposed here and applied to the problem. This method
896: makes use of the analytic properties of apparent and event
897: horizons.
898: 
899: To motivate the method, note that if the event horizon were
900: stationary at $t = 10M$ then the expansion of the surface would be
901: vanishing there:
902: \begin{equation}
903: \theta\left(t = 10M\right) = \frac{1}{A}\frac{dA}{dt} = 0.
904: \end{equation}
905: In such a circumstance the apparent horizon could be used
906: as initial data for the eikonal, which could
907: then be propagated into the past. However, according to the second
908: law of black hole mechanics, in the case that the horizon is 
909: nonstationary, which is the situation expected for this asymmetric
910: problem, at $t = 10M$ the horizon will satisfy
911: \begin{equation}
912: \theta \left(t = 10M\right)\geq 0,
913: \end{equation}
914: in the forward time direction. In the backwards time direction
915: these dynamics correspond to $\theta \leq 0$ at $t = 10M$. As
916: such, the Taylor series of any compact null surface,
917: including the critical surface of the event horizon, is at least
918: of the form
919: \begin{equation}
920: A\left(t + dt\right) = A\left(t\right) + dt A \theta + dt^{2}
921: \frac{d^{2}A}{2dt^{2}} + \ldots
922: \end{equation}
923: and probably cannot be truncated to below
924: \begin{equation}\label{adot}
925: A\left(t + dt\right) \approx A\left(t\right) + dt A \theta +
926: \ldots
927: \end{equation}
928: The apparent horizon is then a poor estimate for the
929: the event horizon at $t = 10M$. The objective is to use this 
930: behavior (\ref{adot}) in combination with the
931: property that outgoing null data followed into the past converges
932: to the event horizon. The hope is to establish a better
933: approximation for the structure of the section of the event
934: horizon at $t = 10M$, which can then be used in an event horizon
935: tracking method.
936: 
937: To proceed, consider three compact null surfaces of outgoing data:
938: One completely interior to the horizon $\Gamma^{i}$, one
939: completely exterior to the horizon $\Gamma^{e}$, and one that is a
940: surface of  section of the event horizon $\Gamma^{h}$. Let
941: $t_{\downarrow} \rightarrow \infty$ denote propagation into the
942: past. By the property that outgoing data numerically converges to
943: the horizon when propagated into the past, $\lim_{t_{\downarrow}
944: \rightarrow \infty} \Gamma^{i} = \lim_{t_{\downarrow} \rightarrow
945: \infty} \Gamma^{e} =  \lim_{t_{\downarrow} \rightarrow \infty}
946: \Gamma^{h}$. Thus in most cases, if the surface
947: $\Gamma^{i}\left(t_{\downarrow} + dt_{\downarrow}\right)$ is
948: pulled back to the time level $t_{\downarrow}$ and compared to the
949: surface $\Gamma^{i}\left(t_{\downarrow} \right)$ it will be
950: completely exterior to $\Gamma^{i}\left(t_{\downarrow} \right)$.
951: Similarly, if the surface $\Gamma^{e}\left(t_{\downarrow} +
952: dt_{\downarrow}\right)$ is pulled back to the time level
953: $t_{\downarrow}$ it will be completely interior to the surface
954: $\Gamma^{e}\left(t_{\downarrow} \right)$.  That is, by iteratively
955: pulling the surfaces back to a single time level, spherically
956: symmetric data will approach the numerical event horizon.
957: Exceptions to this general behavior stem from the presence of
958: caustics in the spacetime, where null surfaces intersect, and in
959: neighborhoods of the event horizon. In particular, as discussed
960: above, the horizon of this asymmetric collision data is expected
961: to satisfy $\theta \geq 0 $. If the
962: surface $\Gamma^{h}\left(t_{\downarrow} + dt_{\downarrow}\right)$
963: is then pulled back to  $t_{\downarrow}$ it will  be interior to
964: the surface $\Gamma^{h}\left(t_{\downarrow}\right)$. Let
965: $\Gamma^{i}_{\epsilon}\left(t_{\downarrow}\right)$ be a
966: perturbation of $\Gamma^{h}\left(t_{\downarrow}\right)$ that is
967: arbitrarily close to the horizon
968: $\Gamma^{h}\left(t_{\downarrow}\right)$ but completely interior to
969: the surface $\Gamma^{h}\left(t_{\downarrow}\right)$. Due to round
970: off and truncation error of the finite difference approximation
971: and the fact that $\lim_{t_{\downarrow}\rightarrow \infty}
972: \Gamma^{i}_{\epsilon} = \Gamma^{h}$, if the numerical finite
973: difference approximation
974: $\hat{\Gamma}^{i}_{\epsilon}\left(t_{\downarrow} +
975: dt_{\downarrow}\right)$, which approximates the continuum section
976: $\Gamma^{i}_{\epsilon}\left(t_{\downarrow} +
977: dt_{\downarrow}\right)$, is pulled back to the time level
978: $t_{\downarrow}$ it will generically intersect
979: $\hat{\Gamma}^{i}_{\epsilon}\left(t_{\downarrow}\right)$ and
980: contain neighborhoods that are both interior and exterior to the
981: finite difference approximation
982: $\hat{\Gamma}^{i}_{\epsilon}\left(t_{\downarrow}\right)$. Similar
983: behavior will hold for numerical data
984: $\hat{\Gamma}^{e}_{\epsilon}\left(t_{\downarrow}\right)$ defined
985: to be a perturbation of
986: $\hat{\Gamma}^{h}\left(t_{\downarrow}\right)$ that is arbitrarily
987: close to the horizon $\hat{\Gamma}^{h}\left(t_{\downarrow}\right)$
988: , and completely exterior to
989: $\hat{\Gamma}^{h}\left(t_{\downarrow}\right)$.
990: 
991: According to this behavior, data for the eikonal at the time level
992: $t = 10M$ can be constructed by iteration over that time slice.
993: This approach is similar to treating the event horizon as if it
994: were an apparent horizon, although modified to account for
995: this nonstationary regime. Beginning with spherically symmetric
996: initial data that is well exterior to the horizon
997: $\hat{\Gamma}^{e}_{1}\left(t_{\downarrow} = 10M\right)$ the data
998: is updated, pulled back to the original time level and reset as
999: follows: $\hat{\Gamma}^{e}_{2}\left(t_{\downarrow}\right) =
1000: \hat{\Gamma}^{e}_{1}\left(t_{\downarrow} +
1001: dt_{\downarrow}\right)$. The step
1002: $\hat{\Gamma}^{e}_{n+1}\left(t_{\downarrow}\right) =
1003: \hat{\Gamma}^{e}_{n}\left(t_{\downarrow} + dt_{\downarrow}\right)$
1004: is then repeated for several hundred iterations, which corresponds
1005: to ${\mathcal{O}}\left(10\right)$ $e$ folding times. Note that in
1006: the case that the spacetime is stationary the surface
1007: $\hat{\Gamma}$ will converge to the apparent horizon according to
1008: this method. Since the apparent horizon of a stationary spacetime
1009: coincides with the event horizon, this method will generically
1010: find the event horizon in the case of stationary spacetimes.
1011: However, in the nonstationary regime, which is the case for the
1012: asymmetric collision problem, the result of this procedure is at
1013: best an improved initial guess for an event horizon tracker.
1014: Further information, such as area analysis or study of the
1015: apparent horizon, is required to argue that the resulting surface
1016: is a candidate section of the event horizon.
1017: 
1018: 
1019: \begin{figure} % Imported eps example.
1020: \epsfxsize=8cm
1021: \centerline{\epsfbox{vImp1.ps}}
1022: \vspace{0.5cm}
1023: \caption{Tanh
1024: data (see Eq(\ref{id}) for the eikonal in asymmetric binary black hole
1025: coalescence. The black contour is the estimate of the event horizon
1026: section, while the wire mesh is the apparent horizon. These data are set at the end 
1027: of the computational evolution ($t=10M$), and will be evolved into the past.} \label{m:1}
1028: \end{figure}
1029: 
1030: We show in figure (\ref{m:1}) the result of such a process, applying
1031: $\approx 200$ iterations over the time slice $t = 10M$. Figures (\ref{m:1})
1032: through (\ref{bh:8}) show the eikonal function in the $z=0$ plane. The location
1033: of the determined guess for the surface $\Gamma$ (which we will take to begin
1034: our evolution of the horizon into the past) is encoded into the eikonal
1035: by the color map. With this $\Gamma$, data for the eikonal can be written
1036: in the form: 
1037: \begin{equation}\label{id}
1038: S\left(0,x^{i}\right) = 1 + \tanh\left(\frac{r_{c} - r}{c}\right)
1039: \end{equation}
1040: In equation (\ref{id}) the first argument of the eikonal is $t_{\downarrow} = 0$;
1041: $t_{\downarrow}$ will increase into the past. Also, $r_{c}$
1042: denotes the data $\Gamma$ and $c$ controls its steepness. Typically
1043: we take a transition width $c$  on the order of a computational zone. Note
1044: that this surface is not considered to be a true section of the
1045: event horizon, but instead is a good initial guess or candidate
1046: section, which after a few $M$ will evolve into better
1047: approximation of a true section of the event horizon. By way of
1048: comparison figure (\ref{m:1}) also shows at $t = 10M$ the final apparent
1049: horizon as a white wire mesh. Note that both the apparent horizon and
1050: these data for the eikonal are highly distorted from the
1051: stationary case. Figure (\ref{m:2}) shows the resulting eikonal
1052: function $S\left(x,y,z\right)$ after $2M$ of evolution into the past.  
1053: Note that
1054: the surface $\Gamma$ is not qualitatively changed during the
1055: evolution.
1056: 
1057: \begin{figure} % Imported eps example.
1058: \epsfxsize=8cm
1059: \centerline{\epsfbox{vImp2.ps}}
1060: \vspace{0.5cm}
1061: \caption{Data of figure (\ref{m:1}) (asymmetric binary black hole
1062: coalescence) evolved backward from $t = 10M $ to $t = 8M$.} \label{m:2}
1063: \end{figure}
1064: 
1065: \section{Surface Extraction and Apparent Horizons}
1066: Figure
1067: %s  (\ref{m:3}) and  
1068: (\ref{m:4}) shows several frames of the
1069: evolution of the eikonal using a viscosity solution of $\epsilon =
1070: h^{2}$ (not of the improved viscosity form). This figure
1071: display the value of the eikonal function on the $z=0$ plane. 
1072: %Figure (\ref{m:3}) shows these values via a color map, which is the
1073: %same as figure (\ref{m:1}). 
1074: Figure (\ref{m:4}) shows the eikonal
1075: data as an elevation map and also via the color map of (\ref{m:1}). 
1076: Note in those figures that null surfaces
1077: interior to the event horizon undergo a change in topology and
1078: this topological transition is continuously monitored by the
1079: viscosity solutions of the eikonal. In figure   (\ref{m:4})
1080: $t = 0.562 M$  is shown in the upper left-hand corner,
1081: $t=1.5M$ is shown in the upper right-hand corner, $t = 2.5M$ is shown 
1082: in the lower
1083: left-hand corner and  $t = 5.0M$ appears in the lower
1084: right-hand corner.
1085: 
1086: 
1087: %\begin{figure} % Imported eps example.
1088: %\epsfxsize=8cm
1089: %\centerline{\epsfbox{lvlst1.ps}}
1090: %\vspace{0.5cm}
1091: %\caption{Evolution
1092: %of the eikonal for asymmetric binary black hole coalescence.}
1093: %\label{m:3}
1094: %\end{figure}
1095: 
1096: \begin{figure} % Imported eps example.
1097: \epsfxsize=8cm
1098: \centerline{\epsfbox{lvlst2.ps}}
1099: \vspace{0.5cm}
1100: \caption{Change of
1101: topology in eikonal for asymmetric binary black hole coalescence, 
1102: shown as an elevation map.}
1103: \label{m:4}
1104: \end{figure}
1105: 
1106: The figures (\ref{m:5}) - (\ref{bh:8}) continue the sequence of 
1107: figures (\ref{m:1}) - (\ref{m:2}), and show, for several
1108: values of $t_{\uparrow}$, the value of the eikonal in the $z=0$ plane;
1109: the location of the apparent horizon in the $3$ dimensions (the white
1110: wire frame); and in the black wire frame, locations of sections of a  candidate event horizon
1111: $\hat{\Gamma}_{c}\left(t_{\downarrow}\right)$ that is generated by
1112: evolution of the eikonal equation from data constructed
1113: using the method described in the previous section. In this
1114: context, the surfaces
1115: $\hat{\Gamma}_{c}\left(t_{\downarrow}\right)$ are extracted from
1116: the eikonal data using the technique described in section IV.
1117: Note that
1118:  $\hat{\Gamma}_{c}$ completely
1119: contains the apparent horizons throughout their evolution. This is
1120: a fundamental condition that any numerically constructed black
1121: hole event horizon must satisfy. To determine how these results
1122: depend on the initial data $\hat{\Gamma}_{c}\left(t_{\downarrow} =
1123: 0 \right)$, choosing initial data
1124: $\hat{\Gamma}_{\delta}\left(t_{\downarrow} = 0 \right)$ of the
1125: form
1126: \begin{equation}
1127: u_{\delta}\left(\theta, \phi\right) = u_{c}\left(\theta,
1128: \phi\right) - \delta,
1129: \end{equation}
1130: permits survey about the data $u_{c}\left(\theta, \phi\right)$,
1131: where $u_{c}\left(\theta, \phi\right)$ corresponds to the data
1132: $\hat{\Gamma}_{c}\left(t_{\downarrow}= 0\right)$. Studies with
1133: $\delta = M/2, M, 2 M $ establish that the level sets
1134: $\hat{\Gamma}_{\delta}$ penetrate both apparent horizons for any
1135: $\delta \geq M/2$. These results suggest that the true event
1136: horizon is contained in a domain parameterized by  $0 < \delta <
1137: M/2 $.
1138: 
1139: \begin{figure} % Imported eps example.
1140: \epsfxsize=8cm
1141: \centerline{\epsfbox{eh2.ps}}
1142: \vspace{0.5cm}
1143: \caption{Asymmetric
1144: binary black hole coalescence:  $t_{\uparrow} = 8M$. This is the 
1145: same as figure (\ref{m:2}) but the black wire mesh shows the estimated 
1146: location of the event horizon.} \label{m:5}
1147: \end{figure}
1148: 
1149: \begin{figure} % Imported eps example.
1150: \epsfxsize=8cm
1151: \centerline{\epsfbox{eh5.ps}}
1152: \vspace{0.5cm}
1153: \caption{Asymmetric
1154: binary black hole coalescence:  $t_{\uparrow} = 5M$. Note 
1155: that while the apparent horizons (the white wire-frame
1156: ``spheres") are still well separate at $t=5M$, 
1157: the event horizon (black wire frame ``peanut" 
1158: already has one component only.} \label{m:6}
1159: \end{figure} 
1160: 
1161: \begin{figure} % Imported eps example.
1162: \epsfxsize=8cm
1163: \centerline{\epsfbox{eh7.ps}}
1164: \vspace{0.5cm}
1165: \caption{Asymmetric
1166: binary black hole coalescence: $t_{\uparrow} = 3M$.} \label{m:7}
1167: \end{figure}
1168: 
1169: \begin{figure} % Imported eps example.
1170: \epsfxsize=8cm
1171: \centerline{\epsfbox{eh9.ps}}
1172: \vspace{0.5cm}
1173: \caption{Asymmetric
1174: binary black hole coalescence: $t_{\uparrow} = 1M$.
1175: Careful inspection of the event horizon (black wireframe) suggests two separated 
1176: components.} \label{bh:8}
1177: \end{figure}
1178: 
1179: Figures  (\ref{m:8}) and  (\ref{m:9}) show two views of the
1180: extracted level set $S = 0$ at $t=2M$. Note that this surface is highly
1181: distorted and shows the event horizon just after merger. Also shown in
1182: these figures are the apparent horizons for the two black holes.
1183: In these figures each color of the color
1184: map denotes a level set $\hat{\Gamma}$ of the eikonal. As such,
1185: each color represents a null surface. From these figures it is
1186: apparent that at this time there appears one innermost null
1187: surface that completely contains both apparent horizons. Thus, the results
1188: of the viscosity solutions suggest a merger time much closer to
1189: $2M $ then the $8M$ found with the apparent horizon trackers.
1190: [Note that in figures  (\ref{m:8}) and  (\ref{m:9}) the event horizon is 
1191: shown as a white wire mesh, while the apparent horizons are shown
1192: as black wire meshes. This is opposite to the color scheme used 
1193: in figures \ref{m:1}, \ref{m:2}, \ref{m:5} - \ref{bh:8}, which was an 
1194: independent study of the evolution as opposed to the study of the throat
1195: geometry considered here.]
1196: \begin{figure} % Imported eps example.
1197: \epsfxsize=8cm
1198: \centerline{\epsfbox{topology1.ps}}
1199: \vspace{0.5cm}
1200: \caption{Level
1201: set extraction for asymmetric binary black hole coalescence: I.}
1202: \label{m:8}
1203: \end{figure}
1204: 
1205: 
1206: \begin{figure} % Imported eps example.
1207: \epsfxsize=8cm
1208: \centerline{\epsfbox{topology2.ps}}
1209: \vspace{0.5cm}
1210: \caption{Level
1211: set extraction for asymmetric binary black hole coalescence: II.}
1212: \label{m:9}
1213: \end{figure}
1214: 
1215: 
1216: \section{Change of Topology}
1217: As shown in figure (\ref{m:4}) the viscosity solutions of the
1218: eikonal equation do continuously monitor a change in topology. In
1219: the case of asymmetric binary black hole coalescence it is
1220: conjectured that the level set $\Gamma$, which corresponds to a
1221: section of the event horizon, must take a higher genus topology at
1222: merger. To investigate this possibility, figure (\ref{m:10}) shows
1223: the level set $\hat{\Gamma}$ viewed along the axis joining the
1224: centers of the apparent horizons. In that figure it is apparent
1225: that the throat function of the topological transition is
1226: elliptical in geometry. Studies indicate that this elliptical
1227: throat function persists for all null surfaces (i.e. those 
1228: slightly inside or slightly outside our candidate event horizon) undergoing the
1229: topological transition. Further, for all of our computed transitions of the
1230: null surfaces, no higher genus topology is exhibited; instead,
1231: the elliptical geometry of the throat persists to the transition.
1232: These results suggest that if there is a non trivial topology in
1233: the sections of the event horizon as a consequence of the
1234: asymmetry of the merger, then that topology change is bounded to occur
1235: when the minor axis of the ellipse is within one of our computational 
1236: zones, or $h = M /4$.
1237: 
1238: \begin{figure} % Imported eps example.
1239: \epsfxsize=8cm
1240: \centerline{\epsfbox{topology3.ps}}
1241: \vspace{0.5cm}
1242: \caption{Throat
1243: function for asymmetric binary black hole coalescence:
1244: $t_{\uparrow} = 1.562 M $.} \label{m:10}
1245: \end{figure}
1246: 
1247: \section{Area Analysis}
1248: \begin{figure} % Imported eps example.
1249: \epsfxsize=8cm
1250: \centerline{\epsfbox{asymmarea.ps}}
1251: \vspace{0.5cm}
1252: \caption{Area
1253: versus time for asymmetric binary black hole coalescence.The horizontal 
1254: scale is $t_{\downarrow}$, i.e. time measured into the past. 
1255: The curves are (bottom to top) for 
1256: $\epsilon = 2 h^{2}$, $\epsilon = h^{2}$, and the improved viscosity
1257: solution.}
1258: \label{n:1}
1259: \end{figure}
1260: 
1261: 
1262: 
1263: \section{Black Hole Areas}
1264: 
1265: Figure  (\ref{n:1}) shows an area versus time plot for this
1266: asymmetric collision. The curve with
1267: the lowest area is the result of a viscosity solution with
1268: $\epsilon = 2 h^2$. The curve with the second lowest area is the
1269: result of a viscosity solution with $\epsilon = h^{2}$, while the
1270: topmost curve is an improved viscosity solution
1271: composed of the two higher viscosity solutions. 
1272: According to these results it is immediately apparent that the
1273: viscosity solutions of higher viscosity show a merger time that is
1274: later in $t_{\uparrow}$ (and therefore prior in $t_{\downarrow}$)
1275: then the merger time found with solutions constructed in  the
1276: limit of vanishing viscosity. These results then indicate that the
1277: error (or bias) in the merger time of the viscosity solutions is
1278: directly related to the magnitude of the viscosity. More
1279: precisely, for the continuum merger time of $t^{*}_{\uparrow}$ and
1280: an approximate merger time of ${t^{*}_{\uparrow}}_{\epsilon}$,
1281: constructed using a viscosity solution of viscosity parameter
1282: $\epsilon$, the function $f\left(\epsilon\right) =
1283: {t^{*}_{\uparrow}}_{\epsilon} - t^{*}_{\uparrow}$ is increasing in
1284: $\epsilon$. (Here all
1285: surface areas are calculated with  $m^{2}$ points where $m=100$.)
1286: 
1287: \begin{figure} % Imported eps example.
1288: \epsfxsize=8cm
1289: \centerline{\epsfbox{conv.ps}}
1290: \vspace{0.5cm}
1291: \caption{Areas
1292: versus time for asymmetric binary black hole coalescence. The horizontal 
1293: scale is $t_{\downarrow}$. }
1294: \label{n:2}
1295: \end{figure}
1296: 
1297: 
1298: 
1299: Figure (\ref{n:2}) shows several area versus time curves for
1300: initial data of the form
1301: \begin{equation}
1302: u_{\delta}\left(\theta, \phi\right) = u_{c}\left(\theta,
1303: \phi\right) - \delta
1304: \end{equation}
1305: where the data $u_{c}\left(\theta, \phi\right)$ is that obtained
1306: using the method of section VII. From top to bottom, the curves show areas for $\delta
1307: = 0, M/4, M/2$ and with a viscosity parameter of $\epsilon =
1308: h^{2}$. Recall that studies of the apparent horizons found that
1309: the true event horizon is contained in the domain $0 < \delta <
1310: M/2$.  This survey over $\delta$ is conducted in search for the
1311: convergence signature associated to event horizons. Due to the time
1312: scale of this data, the time scale of the dynamics, and the
1313: relaxation time scale of the event horizon tracking method, the
1314: signature is not clearly apparent. However, this study of the area
1315: curves does show convergence of the areas, which is expected for
1316: null data approaching the horizon when propagated into the past.
1317: The curve with $\delta = M/2$ shows behavior suggestive of an
1318: event horizon since the $\delta \neq M/2$ curves all approach that of
1319: $\delta =  M/2$. The $\delta =  M/2$ curve is therefore considered
1320: the best candidate for the numerical event horizon of this study.
1321: Note that the sections of this event horizon completely contain
1322: the correct apparent horizons for all $t_{\downarrow} < 9M$.
1323: Further, these $\delta = M/2$ data show a bifurcation time at $t_{\downarrow}
1324: \approx 8.3 M$, which corresponds to a merger time of about
1325: $t_{\uparrow} \approx 1.7 M$. This bifurcation time is detected by
1326: an algorithm that searches for any points $\theta, \phi$ of the
1327: surface such that $u\left(\theta, \phi\right) < h $. In the
1328: circumstance that a point is found such that $u\left(\theta,
1329: \phi\right) < h $, the bifurcation is expected to occur in a few
1330: more $dt = M/16$ in the $t_{\downarrow}$ direction. A few time
1331: levels prior to this bifurcation time in $t_{\downarrow}$ (i.e. just
1332: after the bifurcation in $t$), the single merged horizon component 
1333: has an area (the area of the level set) computed to be
1334: $A = 148.9$. Based on analysis of area computations for exact solutions 
1335: (figure(\ref{vserrora:1}), we anticipate an error in the horizon area of
1336: several percent. We conservatively assign an $8\%$ error to the areas.
1337: Just prior (in $t$) to merger the area of each black hole is
1338: then $A = 74.5 \pm 6.0 M^2$. These individual areas correspond to a
1339: Schwarzschild mass of about $M = 1.48 \pm 0.12$. This result is a
1340: substantially larger mass for each hole then determined by
1341: the apparent horizon finders at the time level $t_{\uparrow} = 0$.
1342: Interestingly, studies have found that apparent horizons separated
1343: by about $10M$ show an increase in their mass
1344: due to the effects of binding energy\cite{bonning}. Individual masses of about
1345: $M \approx 1.36$ are then only a slight departure from studies
1346: that account for the binding energy of the holes. Further, at the
1347: time of merger the holes have undergone $1.7 M$ of evolution,
1348: during which the holes could accrete any surrounding gravitational
1349: radiation present in the initial data. The presence of such
1350: radiation would lead to larger masses then those found using
1351: apparent horizon finders at $t_{\uparrow} = 0$. However, it is
1352: important to note that due to the viscosity in the solution the
1353: result $M \approx 1.36$ can only properly be considered as a lower
1354: bound on the calculated masses. 
1355: The most significant contribution to any error in this result
1356: must stem from the relatively small time scale of this asymmetric
1357: collision data and coupling of that time scale to the $e-$folding
1358: time scale of this event horizon detection method.
1359: 
1360: 
1361: \section{Conclusions}
1362: In this work we have demonstrated a relatively 
1363: simple yet robust and (most importantly) generic solution
1364: to the problem of numerically tracking black hole event horizons.
1365: An implementation of our method made use of an explicit second
1366: order diffusion term to regulate the solution singularities
1367: associated to caustics. As demonstrated by analysis of 
1368: analytic sources, this term does introduce numerical 
1369: error although we demonstrate our control over these effects and 
1370: the resulting accuracy. But, the use of a second order
1371: diffusion term is not required by our method per se; 
1372: and a variety of other approaches can be employed. Examples of 
1373: other methods for controlling breakdown of the numerical solution
1374: include those classes of high resolution shock capturing
1375: numerical schemes that are used extensively in computational 
1376: fluid dynamics for hyperbolic problems similar to the eikonal equation.
1377: 
1378: The application of our new method  for event horizon tracking method
1379: considered the asymmetric binary black hole coalescence problem,
1380: including a detailed analysis of areas of the surfaces of sections,
1381: the collision time, associated apparent horizons, and the topology of the 
1382: horizon. Due
1383: to the relatively short time scale of the collision data, our method
1384: was unable to demonstrate the signature of the black hole
1385: event horizon. We believe that this problem is due to the data 
1386: itself and not due to our method. 
1387: We anticipate much more accurate and convincing results as more
1388: accurate computational simulations of black hole interactions 
1389: become available. 
1390: 
1391: % ({\it REVTEX} 3.0 automatically issues
1392: % a \newpage command when the \begin{table} or \begin{figure}
1393: % commands are used, so the figures and tables will be placed
1394: % on separate pages by {\it REVTEX}).
1395: 
1396: 
1397: 
1398: 
1399: \begin{references}
1400: % Please use the \bibitem command to create references.
1401: \bibitem{he:}
1402:   S. W. Hawking and G. F. R. Ellis,
1403:   ``The Large Scale Structure of Spacetime",
1404:   Cambridge University Press(1973).
1405: 
1406: \bibitem{opsny:}
1407:   J. R. Oppenheimer  and H. Snyder,
1408:   Phys. Rev.,
1409:   {\bf 56},
1410:   455
1411: (1939).
1412: 
1413: 
1414: \bibitem{carter:}
1415:   B. Carter,,
1416:   Phys. Rev. Lett.,
1417:   {\bf 26},
1418:   331
1419:   (1971).
1420: \bibitem{bch:}
1421:   J. M. Bardeen, B.Carter, and S. W. Hawking,
1422:   Commun. Math. Phys.,
1423:   {\bf 31},
1424:   161
1425: (1973).
1426: 
1427: 
1428: \bibitem{wald2:}
1429:   R. M. Wald,
1430:   ``Quantum Field Theory in Curved Spacetime and Black Hole Thermodynamics",
1431:   University of Chicago Press
1432:   (1994).
1433: 
1434: 
1435: \bibitem{choptuik:}
1436:   M. W. Choptuik,
1437:   Phys. Rev. Lett.,
1438:   {\bf 70},
1439:   9
1440:   (1993).
1441: 
1442: 
1443: 
1444: 
1445: 
1446: 
1447: \bibitem{as1:}
1448: A. Celotti, J. C. Miller, and D. W. Sciama,
1449: Class. Quant. Grav.,
1450: {\bf 16},
1451: 3	
1452: (1999).
1453: 
1454: \bibitem{e1:}
1455: K. Menou, E. Quataert, and R. Narayan,
1456: Proc. $8^{th}$ Marcel Grossmann Meeting on General Relavity,
1457: (1997).
1458: 
1459: \bibitem{ambh:}
1460: Richstone \textit{et} \textit{al},
1461: Nature,
1462: {\bf 395},
1463: A14
1464: (1998).
1465: 
1466: \bibitem{eb:}
1467: K. Belczynski, V. Kalogera, and T. Bulik,
1468: Astrophys. J.,
1469: {\bf 572},
1470: 407
1471: (2001).
1472: 
1473: \bibitem{lehner:}
1474: L. Lehner,
1475: Class. Quantum. Grav.,
1476: {\bf 18},
1477: 161
1478: (2001).
1479: 
1480: 
1481: \bibitem{bbhc1:}
1482: L. P. Grischuk \textit{et} \textit{al},
1483: Physics-Uspekhi,
1484: {\bf 171},
1485: 3
1486: (2001).
1487: 
1488: 
1489: \bibitem{eh1:}
1490: S. A. Hughes \textit{et} \textit{al},
1491: Phys. Rev D.,
1492: {\bf 49},
1493: 4004
1494: (1994).
1495: 
1496: \bibitem{eh2:}
1497: P. Anninos \textit{et} \textit{al},
1498: Phys. Rev Lett.,
1499: {\bf 74},
1500: 630
1501: (1995).
1502: 
1503: 
1504: \bibitem{eh3:}
1505: R. A. Matzner \textit{et} \textit{al},
1506: Science,
1507: {\bf 270},
1508: 941
1509: (1995).
1510: 
1511: \bibitem{eh4:}
1512: J. Libson \textit{et} \textit{al},
1513: Phys. Rev. D.,
1514: {\bf 53},
1515: 4335
1516: (1996).
1517: 
1518: \bibitem{eh5:}
1519: J. Masso \textit{et} \textit{al},
1520: Phys. Rev. D.,
1521: {\bf 59},
1522: (1999).
1523: 
1524: \bibitem{is1:}
1525: A. Ashtekar \textit{et} \textit{al},
1526: Phys. Rev. Lett.,
1527: {\bf 85},
1528: 3564
1529: (2000).
1530: 
1531: \bibitem{newman1:}
1532: J. Ehlers and E. T. Newman,
1533: J. Math. Phys.,
1534: {\bf 41},
1535: 3344
1536: (2000).
1537: 
1538: \bibitem{newman2:}
1539: S. Frittelli and E. T. Newman,
1540: J. Math. Phys.,
1541: {\bf 40},
1542: 383
1543: (1999).
1544: 
1545: \bibitem{newman3:}
1546: E. T. Newman and A. Perez,
1547: J. Math. Phys.,
1548: {\bf 40},
1549: 1093
1550: (1999).
1551: 
1552: \bibitem{icn:}
1553: S. A. Teukolsky,
1554: Phys. Rev. D,
1555: {\bf 61},
1556: 087501
1557: (2000).
1558: 
1559: \bibitem{wini2:}
1560: S. Husa and J. Winicour,
1561: Phys. Rev D.,
1562: {\bf 60},
1563: 84019
1564: (1999).
1565: 
1566: \bibitem{exc1:}
1567: S. Brandt \textit{et} \textit{al},
1568: Phys. Rev. Lett.,
1569: {\bf 85},
1570: 5496
1571: (2000).
1572: 
1573: 
1574: 
1575: \bibitem{deirdre:}
1576: D. M. Shoemaker, M. F. Huq, and R. A. Matzner,
1577: Phys. Rev. D,
1578: {\bf 62},
1579: 124005
1580: (2000).
1581: 
1582: \bibitem{bonning}E. Bonning, D. Neilsen, and Richard A. Matzner,
1583: "Physics an dInitial Data for Black Hole Spacetimes", in preparation
1584: (2003).
1585: 
1586: \end{references}
1587: 
1588: 
1589: \end{document}
1590: 
1591: % end of file Template.tex
1592: