gr-qc0307055/ce.tex
1: %\documentclass[prd, twocolumn, aps, showpacs]{revtex4}
2: \documentclass[prd, preprint, aps, showpacs]{revtex4}
3: 
4: \usepackage{graphicx}
5: \usepackage{color}
6: 
7: \def\Lie{\hbox{\it\char'44}}
8: \def\be{\begin{equation}}
9: \def\ee{\end{equation}}
10: \def\ba{\begin{eqnarray}}
11: \def\ea{\end{eqnarray}}
12: \def\beq{\begin{eqnarray}}
13: \def\eeq{\end{eqnarray}  }
14: 
15: \def\eref#1{Eq.~(\ref{#1})}
16: \def\Eref#1{Eq.~(\ref{#1})}
17: \def\fref#1{Fig.~\ref{#1}}
18: \def\Fref#1{Fig.~\ref{#1}}
19: \def\sref#1{Section~\ref{#1}}
20: \def\Sref#1{Section~\ref{#1}}
21: \def\tref#1{Table~\ref{#1}}
22: \def\Tref#1{Table~\ref{#1}}
23: 
24: \def\Comment#1{{\textcolor{red}{\bf #1}}}
25: \def\BlueComment#1{{\textcolor{blue}{\bf #1}}}
26: 
27: \def\rmd{{\rm d}}
28: \def\rmO{{\rm O}}
29: \def\ADM{{\rm ADM}}
30: \def\HOR{{\rm H}}
31: \def\ISCO{{\rm ISCO}}
32: \def\BE{{E_b}}
33: \def\lp{\left(}
34: \def\rp{\right)}
35: \def\lb{\left[}
36: \def\rb{\right]}
37: \def\etal{{et al.}}
38: 
39: 
40: 
41: \begin{document}
42: \title{Extended Lifetime in Computational
43: Evolution of Isolated Black Holes}
44: 
45: \author{Matthew Anderson}
46: \author{Richard A. Matzner}
47: \affiliation{Center for Relativity, University of Texas at Austin,
48: Austin, TX 78712-1081, USA}
49: 
50: \begin{abstract} Solving the 4-d Einstein equations as evolution
51: in time requires solving equations of two types: the four elliptic initial
52: data (constraint) equations, followed by the six second order evolution
53: equations. Analytically the constraint equations remain solved under the
54: action of the evolution, and one approach is to simply monitor them ({\it
55: unconstrained} evolution).
56: 
57: The problem of the 3-d computational simulation of even a single
58: isolated vacuum black hole has proven to be remarkably difficult.
59: Recently, we have become aware of two publications that describe
60: very long term evolution, at least for single isolated black
61: holes.
62: An essential feature in each of these results is {\it constraint
63: subtraction}. Additionally, each of these approaches is based on what we
64: call ``modern," hyperbolic
65: formulations of the Einstein equations. It is generally assumed, based on
66: computational experience, that the use of such modern formulations is
67: essential for long-term black hole stability. We report here on comparable
68: lifetime results based on the much simpler (``traditional") $\dot g$ - $\dot
69: K$ formulation.
70: 
71:  
72: With specific subtraction of constraints, with a simple
73: analytic gauge, with very simple
74: boundary conditions, and for moderately
75: large domains with moderately fine resolution, we find computational
76: evolutions of isolated nonspinning black holes for times exceeding $1000
77: GM/c^2$.
78: 
79: We have also carried out a
80: series of {\it constrained } 3-d evolutions of single isolated
81: black holes. We find that constraint solution can produce
82: substantially stabilized long-term single hole evolutions.
83: However, we have found that for large domains, neither constraint-subtracted
84: nor constrained  $\dot g$ - $\dot K$
85: evolutions carried out in Cartesian coordinates admit arbitrarily
86: long-lived simulations. The failure appears to arise from features at the
87: inner excision boundary; the behavior does generally improve with
88: resolution.
89: 
90: 
91: \end{abstract}
92: 
93: \pacs{ }
94: 
95: \maketitle
96: 
97: 
98: %----------------------------------------------------------------------
99: %
100: %
101: %
102: %----------------------------------------------------------------------
103: \section{Introduction}
104: \label{sec:intro}
105: 
106: 
107: Binary black hole systems are expected to be the strongest possible
108: astrophysical gravitational wave sources. In the final moments of
109: stellar mass black hole inspiral, the radiation will be detectable in
110: the current (LIGO-class) detectors. If the total binary mass is of the
111: order of $10M_{\odot}$, the moment of final plunge to coalescence will
112: emit a signal detectable by the current generation of detectors from
113: very distant (Gpc) sources. The merger of supermassive black holes in
114: the center of galaxies will be the dominant signal in the
115: spaceborne LISA detector, and detectable out to large redshift.
116: 
117: 
118: Simulation of these mergers will play an important part in the
119: prediction, detection, and the analysis of their gravitational signals
120: in gravitational wave detectors. To do so requires a correct formalism which
121: does not generate spurious singularities during the attempted simulation.
122: Recent important work\cite{schKidteuk}\cite{yo} has been
123: done in extending the computational lifetime
124: of single isolated black hole simulations. We report here on such an
125: extension
126: which demonstrates that {\it constraint subtraction} by itself is adequate
127: to produce very long-lived simulations. We demonstrate this even for a very
128: simple (``traditional") $\dot g$ - $\dot
129: K$ formulation, with specific subtraction of constraints (which
130: are analytically zero) for single isolated nonspinning
131: black holes, with a simple
132: analytic gauge (lapse and shift {\it not} ``densitized";
133: see \ref{sec:densitized} below), with very simple
134: boundary conditions, and for moderately
135: large domains with moderately fine resolution.
136: 
137: We have also found that {\it constrained }3-d evolution {\it with}
138: densitized lapse can produce
139: substantially stabilized long-term single holes, even for subtractions
140: that
141: differ from the values that we have found to be optimal in the
142: unconstrained
143: case, and we give some preliminary constrained evolution results.
144: 
145: However, in all cases we find that attempting to carry out the simulations
146: on very large domains ($\pm 20M$, or larger) still yields eventually
147: unstable simulations, by any of the methods reported here.
148: 
149: 
150: 
151: \section{$3+1$ Formulation of Einstein Equations}
152: 
153: We take a Cauchy formulation
154: (3+1) of the ADM type, after Arnowitt, Deser, and
155: Misner~\cite{ADM}. In such a method the 3-metric $g_{ij}$ and its momentum
156: $K_{ij}$ are specified at one initial time on
157: a spacelike hypersurface, and evolved into the future.  The ADM metric is
158: \be
159: \rmd s^2 = -(\alpha^2 - \beta_i \beta^i)\,\rmd t^2 + 2\beta_i \, \rmd t
160: \,\rmd x^i
161:      + g_{ij}\, \rmd x^i\, \rmd x^j
162: \label{eq:admMetric}
163: \ee
164: where $\alpha$ is the lapse function and $\beta^i$ is the shift
165: 3-vector; these gauge
166: functions encode the
167: coordinatization.\renewcommand{\thefootnote}{\fnsymbol{footnote}}\setcounter
168: {footnote}{2}\fnsymbol{footnote}
169: 
170: \footnotetext[2]{Latin indices run $1,2,3$ and are lowered and raised
171: by $ g_{ij}$
172: and its 3-d inverse $ g^{ij}$.}
173: %\footnote{Latin indices run $1,2,3$ and are lowered and raised by $ g_{ij}$
174: %and its 3-d inverse $ g^{ij}$.}
175: 
176: The Einstein field equations contain both hyperbolic evolution equations
177: and elliptic constraint equations.
178: The constraint equations for vacuum in the ADM decomposition are:
179: \beq
180: H = \frac{1}{2} [R - K_{ij}K^{ij} + K^2] &=& 0,
181: \label{eq:constraintH}
182: \eeq
183: \beq
184: H^i = \nabla_j \lp K^{ij} - g^{ij}K\rp  &=& 0.
185: \label{eq:constraintK}
186: \eeq
187: Here $R$ is the 3-d Ricci scalar constructed from the 3-metric, and
188: $\nabla_j$ is the
189: 3-d covariant derivative compatible with $ g_{ij}$.
190: Initial data must satisfy these constraint
191: equations; one may not freely specify all components of $g_{ij}$ and
192: $K_{ij}$.   
193: 
194: In this paper we are concerned only with single isolated black holes.
195: From this point of view the problem is not
196: solving the initial value equations, since the data are known analytically.
197: Instead, the question is one of the stability of the solution
198: as these data are evolved computationally.
199: The evolution equations from the
200: Einstein system are
201: 
202: \begin{equation}
203:        \dot g_{ij} = -2\alpha K_{ij} +\nabla_j\beta_{i} + \nabla_i\beta_{j}
204: \label{eq:gdot}
205: \end{equation}
206: and 
207: 
208: \begin{equation}
209:       \dot  K_{ij} =  - \nabla_i \nabla_j \alpha + \alpha ( R_{i j}
210:           -2K_{i k}K^k_{~j} + KK_{i j} ) + \beta^k \nabla_k K_{ij} + K_{ik} \nabla_j \beta^k + K_{jk} \nabla_i \beta^k
211: \label{eq:kdot}
212: \end{equation}
213: where a dot ( $\dot{}$ ) denotes the partial derivative with respect
214: to time, and $R_{ij}$ is the 3-d Ricci tensor.
215: 
216: 
217: We call this form of the Einstein equations {\it of ADM type},
218: referring to the fundamental development \cite{ADM}; this specific form is
219: called the $\dot g$ - $\dot K$ form. Here,
220: \eref{eq:constraintH}--\eref{eq:constraintK}, the
221: constraint equations, are the vacuum Einstein equations ${}^4 G_{00} = 0$
222: and
223: ${}^4 G_{0i} = 0$ respectively. \eref{eq:gdot}--\eref{eq:kdot}, the
224: evolution equations, are a first order form of the vacuum Einstein equations
225: ${}^4 R_{ij} = 0$.
226: 
227: The true ADM form writes the evolution equations as
228: ${}^4 G_{ij} = 0$,
229: the space components of the 4-d Einstein tensor, rather than the
230: Ricci tensor. Frittelli and Reula\cite{49Frittelli} have shown that with
231: certain
232: (rather strong) assumptions, there is stable maintenance of the constraints
233: under unconstrained evolution for \eref{eq:gdot}--\eref{eq:kdot},
234: but only neutral stability for ${}^4 G_{ij} = 0$.
235: 
236: 
237: \section{Data Form}
238: \label{sec:DataForm}
239: 
240: In this paper we consider only single isolated black holes,
241: so the data setting problem is already solved; we use Kerr-Schild data,
242: which describes a single isolated spinning or nonspinning black hole
243: hole.
244: 
245: The Kerr-Schild~\cite{KerrSchild} form of a black hole solution describes
246: the 
247: spacetime of a single black hole
248: with mass, $m$, and specific angular momentum, $a = j/m$, in a coordinate
249: system that is well behaved at the horizon:
250: \be
251:         \rmd s^{2} = \eta_{\mu \nu}\,\rmd x^{\mu}\, \rmd x^{\nu}
252:                  + 2H_{KS}(x^{\alpha}) l_{\mu} l_{\nu}\,\rmd x^{\mu}\,\rmd
253: x^{\nu},
254:         \label{eq:1}
255: \ee
256: where $\eta_{\mu \nu}$ is the metric of flat space, $H_{KS}$ is a scalar
257: function of $x^\mu$, and $l_{\mu}$ is an (ingoing) null vector, null
258: with respect to both the background and the full metric,
259: \be
260: \eta^{\mu \nu} l_{\mu} l_{\nu} = g^{\mu \nu} l_{\mu} l_{\nu} = 0.
261: \label{eq:2}
262: \ee
263: 
264: 
265: Comparing the Kerr-Schild metric with the ADM
266: decomposition~\eref{eq:admMetric}, we find that the $t=\hbox{\rm constant}$
267: 3-space metric is: $g_{ij} = \delta_{ij} + 2 H_{KS} l_i l_j$.
268: Further, the ADM gauge variables  are
269: \be
270: \beta_i = 2 H_{KS} l_0 l_i,
271: \ee
272:  and
273: \be
274: \alpha = \frac{1}{\sqrt{1 + 2 H_{KS} l_0^2}}.
275: \label{eq:ksAlpha}
276: \ee
277: 
278: The extrinsic curvature can be computed from Eq.(\ref{eq:gdot}):
279: 
280: \be
281:         K_{ij} = \frac{1}{2\alpha}[\nabla_j\beta_{i} + \nabla_i\beta_{j}
282:                      - \dot g_{ij}],
283: \label{eq:k_ks}
284: \ee
285: 
286: 
287: Each term on the right hand side of this equation is known analytically; in
288: particular, for a black hole at rest, $\dot g_{ij}=0$.
289: 
290: The general non-moving
291: black hole metric in Kerr-Schild form (written in Kerr's original
292: rectangular coordinates) has
293: \begin{equation}
294:         H_{KS} = \frac{mr}{r^{2} + a^{2}\cos^{2} \theta},
295:         \label{eq:ks_h}
296: \end{equation}
297: and
298: \begin{equation}
299:         l_{\mu} = \left(1, \frac{rx + ay}{r^{2} + a^{2}}, \frac{ry -
300:         ax}{r^{2} + a^{2}}, \frac{z}{r}\right),
301:         \label{eq:4}
302: \end{equation}
303: where $r,~ \theta$ (and $\phi$)
304: are auxiliary spheroidal coordinates,  $z = r(x,y,z) \cos \theta$,
305: and $\phi$ is 
306: the axial angle.  $r(x, y, z)$ is obtained from the relation,
307: \begin{equation}
308:         \frac{x^{2} + y^{2}}{r^{2} + a^{2}} + \frac{z^{2}}{r^{2}} = 1,
309:         \label{eq:5}
310: \end{equation}
311: giving
312: \begin{equation}
313:         r^{2} = \frac{1}{2}(\rho^{2} - a^{2}) +
314:         \sqrt{\frac{1}{4}(\rho^{2} - a^{2})^{2} + a^{2}z^{2}},
315:         \label{eq:6}
316: \end{equation}
317: with
318: \begin{equation}
319: \rho = \sqrt{x^{2} + y^{2} + z^{2}}.
320:         \label{eq:rho_def}
321: \end{equation}
322: 
323: 
324: In the
325: nonspinning case, one has $l_i= x_i/r$, so that
326: $\alpha = \frac{1}{\sqrt{1 + 2 m/r}}$, and $\beta_i = 2 m x_i/r$.
327: 
328: 
329: %\begin{figure}
330: %\begin{center}
331: %\includegraphics[width=3.5in,angle=270]{ham.pdf}
332: %\end{center}
333: %\caption{The Hamiltonian constraint (units $m^{-2}$) calculated for .}
334: %\label{fig:hc_with_att}
335: %\end{figure}
336: 
337: 
338: \section{Constraint Subtraction}
339: \label{sec:C-S}
340: 
341: 
342: The difference between \eref{eq:gdot}--\eref{eq:kdot},
343: and ${}^4 G_{ij} = 0$, is a specific subtraction of
344: the constraint equations. This has led to the
345: consideration by a number of groups, of constraint subtraction with
346: coefficients chosen by numerical search, or by analytical estimate (perhaps
347: combined with numerical search) to improve the long-term stability of the
348: unconstrained evolution. We have carried out such a numerical search, and
349: we use the following constraint subtraction:
350: 
351: \be
352: -\alpha H (0.464\, g_{i\,j}\, + 0.36\ K_{i\,j}\,)
353: \label{eq:constraintSub}
354: \ee
355: on the right hand side of the $\dot K_{ij}$ equation (\eref{eq:kdot}).
356: We have found that this subtraction
357: substantially improves the unconstrained evolution
358: of nonspinning single-hole data. For these evolutions we choose fixed
359: (Dirichlet) outer
360: boundary conditions set equal to the analytical value.  In every case the
361: simulation excises the interior of the black hole.
362: One-sided differencing is used near the inner (mask) boundary, so no
363: boundary condition is needed there, consistent with the property of the
364: horizon as a causal boundary. The mask is specified at a radius of 0.75 M.
365: The
366: solution employs the SUNDIALS package for time integration \cite{SUNDIALS,CVODES}.
367: The spatial
368: discretization is fourth order, and the
369: time evolution is typically fourth order (variable-order according
370: to the relative and absolute error
371: tolerances).  This approach successfully stabilizes the evolutions for
372: domains of $\pm 10M$, as shown in Figure \ref{fig:standard_domain}.
373: For larger domains, however, the system is increasingly shorter lived, as shown
374: in Figure \ref{fig:vary_domain}.
375: As a simple measure of the quality of solution, we present the either the
376: $l_2$ norm or rms norm of the Hamiltonian constraint constructed via a
377: straightforward fourth order differencing scheme from the code results.
378: 
379: Although it is difficult to compare subtraction techniques across
380: different formal representations of the Einstein equations, we do find
381: typically much smaller coefficients of subtraction (of order $0.5$) than
382: found for different formulations, e.g. \cite{schKidteuk} with a constraint
383: subtraction of order $-12.$ The BSSN formulation of \cite{yo} is also
384: substantially different from ours (there is a subtraction from the $\dot
385: g_{ij}$
386: equation, for instance, which we do not have, and the subtraction from the
387: $\dot K_{ij}$ equation is different from ours), though the coefficient of
388: subtraction from $\dot K_{ij}$ for this approach is small, comparable to
389: ours.
390: \bigskip
391: 
392: \subsection{Densitized Lapse}
393: \label{sec:densitized}
394: 
395: There is extensive evidence in the literature that a {\it densitized} lapse
396: improves the hyperbolicity\cite{gr-qc/0402123} of the (at best)
397: weakly hyperbolic ADM form of
398: the Einstein equations. We implement densitized lapse for single black hole
399: simulations by writing
400: 
401: \be
402: \alpha = \alpha_{\rm analytic} (g/g_{\rm analytic})^p,
403: \ee
404: where $\alpha_{\rm analytic}$ is the explicit lapse as a function of coordinates
405: given by \eref{eq:ksAlpha},
406: $g_{\rm analytic}$ is the analytic Kerr-Schild $3-$metric determinant
407: as a function of coordinates, ($g_{\rm analytic}=1+2H l_t^2$,
408: \cite{gr-qc/0002076}),
409: $g$ is the computational $3-$metric determinant,
410: and $p$ is an adjustable positive constant usually taken to be $\frac{1}{3}$
411: or
412: $\frac{1}{2}$. In fact we find that densitized lapse does not enhance
413: constraint-subtracted lifetime, though it does contribute substantially
414: to longer lifetime in {\it constrained} evolutions described below.
415: 
416: %{\bf insert  figures here}
417: \begin{figure}
418: \begin{center}
419: \includegraphics[width=5.0in, angle=0]{./figure1}
420: \end{center}
421: \caption{The log of the rms norm of the Hamiltonian
422: constraint violations for
423: constraint-subtracted and unsubtracted nonspinning black hole simulations
424: with excision.  The simulations were performed at resolutions
425: of $M/5$ and $M/7.5$ on a domain size of $\pm 10M$.  The long-lived runs
426: employed optimal constraint subtraction (\eref{eq:constraintSub}). The
427: short-lived run
428: employed no subtraction.
429: }
430: \label{fig:standard_domain}
431: \end{figure}
432: 
433: \begin{figure}
434: \begin{center}
435: \includegraphics[width=5.0in, angle=0]{./figure2}
436: \end{center}
437: \caption{The $l_2$ norm of the Hamiltonian constraint
438: violation for 
439: constraint-subtracted and unsubtracted nonspinning black hole
440: simulations with excision
441: performed at a resolution
442: of $M/5$ on a domain sizes of $\pm 10M$, $\pm 15M$, or $\pm 20M$. All cases
443: employed the optimal constraint subtraction (\eref{eq:constraintSub}).
444: As the computational domain size increases, the simulation is increasingly shorter lived.
445: }
446: \label{fig:vary_domain}
447: \end{figure}
448: 
449: \section{Constrained Evolution}
450: 
451: The evolution of Kerr-Schild data
452: must continue to satisfy the constraint equations,
453: Eqs.~(\ref{eq:constraintH})--(\ref{eq:constraintK})
454: as we evolve away from the initial data.  However,
455: if we choose nonoptimal (e.g. zero)
456: constraint subtraction for even nonspinning black holes, the evolution
457: leads to an eventual violation of the constraints. Hence we have
458: investigated constrained evolution, solving the constraint equations as part
459: of the time update of the evolution equations.
460: 
461: The postevaluated tracking of constraint errors [residual
462: postevaluation] shown in Figures
463: \ref{fig:standard_domain}--\ref{fig:vary_domain} used the direct
464: discretization of the constraint
465: equations \eref{eq:constraintH} and \eref{eq:constraintK}.
466: For constrained evolution we need instead to implement an accurate,
467: efficient method of constraint {\it solution}.
468: We adopt the conformal transverse-traceless method
469: of York and collaborators~\cite{YP}-\cite{Bowen+York} which consists of a
470: conformal decomposition with a scalar $\phi$ that adjusts the metric,
471: and a vector potential $w^i$ that adjusts the longitudinal components of the
472: extrinsic curvature.
473: The constraint equations are then solved for these new quantities  $\phi$,
474: $w^i$ such that
475: the complete solution fully satisfies the constraints.
476: 
477: Applying this approach to constrained evolution,
478: the metric and traceless extrinsic curvature in the middle of a timestep
479: (after an explicit integration forward in time)  are
480: taken as conformal trial functions  $\tilde{g}_{ij}$ and
481: $\tilde{A}^{ij}$.
482: 
483: The physical metric at the end of the full timestep
484: (i.e. after the constraint equation solve), $g_{ij}$,
485: and the trace-free part of the extrinsic
486: curvature at the end of the full timestep, $A_{ij}$,
487: are related to the background fields through a
488: conformal
489: factor:
490: \ba
491: g_{ij} &=& \phi^{4} \tilde{g}_{ij}, \label{confg1} \\
492: \label{confg}
493: A^{ij} &=& \phi^{-10} (\tilde{A}^{ij} + \tilde{(lw)}^{ij}).
494: \label{eq:conf_field}
495: \ea
496: Here $\phi$ is the conformal factor, and $\tilde{(lw)}^{ij}$
497: will be used to cancel any possible longitudinal contribution.
498: $w^i$ is a vector potential, and
499: \ba
500: \tilde{(lw)}^{ij} \equiv \tilde{\nabla}^{i} w^{j} + \tilde{\nabla}^{j} w^{i}
501:         - \frac{2}{3} \tilde{g}^{ij} \tilde{\nabla_{k}} w^{k}.
502: \label{lw}
503: \ea
504: The trace $K$ is not corrected:
505: \be
506: K = \tilde K.
507: \label{tk}
508: \ee
509: Writing the Hamiltonian and momentum constraint equations in terms of
510: the quantities in 
511: Eqs.~(\ref{confg1})--(\ref{tk}), we obtain four coupled
512: elliptic equations for the fields $\phi$ and $w^i$~\cite{YP}:
513: %\ba
514: %\tilde{\nabla}^2 \phi &=&  (1/8) \big( \tilde{R}\phi
515: %        + \frac{2}{3} \tilde{K}^{2}\phi^{5} -   \nonumber \\
516: %        & & \phi^{-7} (\tilde{A}{^{ij}} + (\tilde{lw})^{ij})
517: %            (\tilde{A}_{ij} + (\tilde{lw})_{ij}) \big),   \\
518: %\label{ell_eqPhi}
519: %\tilde{\nabla}_{j}(\tilde{lw})^{ij} &=& \frac{2}{3} \tilde{g}^{ij} \phi^{6}
520: %        \tilde{\nabla}_{j} K - \tilde{\nabla}_{j} \tilde{A}{^{ij}}.
521: %\label{ell_eqs}
522: %\ea
523: \begin{eqnarray}
524: \tilde{\nabla}^2 \phi &=&  \left(1/8\right) \left[ \tilde{R}\phi  \right.
525:        + \frac{2}{3} \tilde{K}^{2}\phi^{5} -   \nonumber \\
526:    & & \phi^{-7} \left(\tilde{A}{^{ij}} + (\tilde{lw})^{ij}\right)
527:      \left.  \left(\tilde{A}_{ij} + (\tilde{lw})_{ij} \right) \right], \label{ell_eqPhi} \\
528:      \tilde{\nabla}_{j}(\tilde{lw})^{ij} &=& \frac{2}{3} \tilde{g}^{ij} \phi^{6}
529:      \tilde{\nabla}_{j} \tilde{K} - \tilde{\nabla}_{j} \tilde{A}{^{ij}}. \label{ell_eqs}
530: \end{eqnarray}
531: 
532: These equations are solved to complete each time-update step.
533: The resulting solved $g_{ij}$ and $K_{ij}$ are taken as the data for the
534: next time-update. Notice that these equations require no specific gauge
535: choice. A similar approach also can be applied to other formulations which
536: generally have a larger number of constraints.
537: 
538: \subsection{Elliptic Equation Boundary Conditions}
539: \label{sec:boundary}
540: 
541: A solution of the elliptic constraint equations requires that boundary
542: data be specified on both the outer boundary {\em and } on the surfaces
543: of any masked regions.
544: For the elliptic solution here we can choose simple
545: conditions, $\phi = 1$ and $w^{i} = 0$, on the masked region surrounding
546: the singularity. Because we solve the problem
547: on a finite domain, we also must provide an outer boundary condition for
548: $\phi $ and $w^{i}$. For this demonstration of the technique, we
549: choose the same conditions at the outer boundary of the domain:
550: $\phi = 1$ and $w^{i} = 0$. In
551: long term evolution we expect the evolved solution to
552: converge (as the solution is refined) to a solution of the constraints,
553: so a global solution $\phi = 1$ and $w^{i} = 0$ is expected in this
554: analytic limit. For achievable resolutions, however,
555: the quantities $\phi $ and $w^{i}$ deviate from this prediction.
556: 
557: 
558: \section{Constrained Evolution Results}
559: \label{sec:physresults}
560: 
561: The elliptic constraint equations are solved either by a
562: PETSc \cite{PETSC_1,PETSC_2,PETSC_3} GMRES solver or
563: KINSOL \cite{SUNDIALS} GMRES solver;
564: the spatial differencing is fourth order. We present
565: below (Figures \ref{fig:constrained} - \ref{fig20.pdf})
566: the results of preliminary constrained evolution of
567: nonspinning black holes. Figure \ref{fig:constrained} shows the
568: rms norm of Hamiltonian and momentum constraints
569: for these simulations.
570: Compared 
571: to the relatively short term crash of the unconstrained
572: evolution, the constrained evolution clearly
573: does
574: stabilize single black hole evolutions in small domains, regardless of the
575: precise
576: subtraction. To fully understand the content of
577: Figure \ref{fig:constrained},
578: consider that, even
579: if the residual limit
580: in the solution of the constraint equations,
581: \eref{ell_eqPhi} -- \eref{ell_eqs}, is set
582: to extremely small values (it can be set very near to machine
583: precision, 
584: meaning that the discretized matrix form of equations
585: \eref{ell_eqPhi} --  \eref{ell_eqs}
586: can be solved to fractional errors of order $10^{-15}$), the
587: post-evaluations 
588: of the constraint residuals will typically show the
589: expected $zero$ only
590: to the internal discretization accuracy (here fourth order).
591: This is why 
592: the constrained solution shown in Figure \ref{fig:constrained}
593: shows a finite (but convergent) level of Hamiltonian
594: constraint violation.
595: 
596: Figures \ref{fig15.pdf}-\ref{fig20.pdf} show a constrained
597: $\dot g - \dot K$ evolution with densitized lapse, at three resolutions
598: ($M/5, M/7.5, M/10$). The $M/5$ evolution became unstable before
599: $t= 100 M$. The $M/7.5$ run began showing large residuals at
600: $t \approx 350 M$. The $M/10$ run shows better behavior than $M/7.5$,
601: at least initially It shows a similar late time instability which
602: tracks (at smaller error) the behavior of the $M/7.5 $ case,
603: but around $150M$ ceases to be convergent; see Figure \ref{fig17.pdf}.
604: Figure  \ref{fig20.pdf} shows the 2d $z=0$ behavior of the residual
605: component 
606: $G_{xx}$ at $t = 100M$. The ``red-blue" pattern of the features
607: near the excision mask indicate that most of the error develops
608: there. The residual becomes more asymmetrical at
609: later times.
610:  
611: \section{CORRECTNESS OF CONSTRAINED EVOLUTION}
612: \label{sec:correctness}
613: 
614: The constraint maintenance approach uses what has been called in
615: magnetohydrodynamics, a {\it projection} method\cite{projection}.  This
616: method for constrained evolution raises questions about the meaning of
617: the solutions obtained. This is sometimes put bluntly: ``Accepting that
618: the method finds solutions of the full Einstein system, how do we know
619: that the found solution is the {\it right} one?" By this is meant that
620: the constraint solution step may somehow move the solution back to an
621: ``erroneous" point on the space of constraint solutions. For instance,
622: it might be possible that although the evolution substep and the
623: constraint substep are individually convergent computational
624: processes, the result of combining them is in
625: some manner {\it not} convergent. (In the much
626: simpler MHD case there are analytical proofs that projection
627: minimizes the resultant error in the magnetic field, in a
628: convergent
629: way.\cite{projection})
630: 
631: \begin{figure}
632: \begin{center}
633: \includegraphics[width=6.0in, angle=0]{./figure3}
634: \end{center}
635: \caption{The rms norm of the Hamiltonian and momentum constraints
636: for simulations of a Schwarzschild (nonspinning)
637: black hole with excision. None of the simulations used any constraint subtraction.
638: The simulations were performed at resolutions 
639: of either $M/5$ or $M/7.5$ on a domain size of $\pm 5M$. The long-lived runs employed
640: constrained evolution as described in the text. The short-lived runs were
641: unconstrained.
642: }
643: \label{fig:constrained}
644: \end{figure}
645: 
646: There are several parts to the response. (It will be clear that we do
647: not pretend 
648: to a rigorous analytical proof.)
649: To begin with, we have
650: constructed completely independent ``residual evaluators"
651: for the full Einstein system\cite{choptuikSB}. These evaluate the
652: Einstein tensor, working just from the metric produced by the
653: computational solution. They are completely different from the way
654: the equations are expressed in the constrained evolution code.
655: As we show in Figures \ref{fig15.pdf},\ref{fig17.pdf}, the
656: resulting residual is in every case initially small (order of
657: truncation error) and convergent. Thus we have
658: achieved a computational solution to the Einstein system.
659: Note that our
660: full Einstein equation residual evaluator
661: checks both the constraints (Einstein equations at one time),
662: and the evolution 
663: equations connecting different time steps.
664: 
665: The residual evaluators are written to return fourth order accurate
666: results. Since they have now been verified and show
667: convergence of the solution, we appeal
668: to the assumption that
669: the Einstein system is not {\it singular} at the solution
670: manifold. Thus we
671: expect that the computational result converges to the analytical
672: solution of the Einstein equations.  We have converged to a
673: spacetime configuration. Physical consistency and generality
674: imply that
675: it is the physically unique one that contains the initial
676: data slice. We also note that, as in the situation in Figure \ref{fig17.pdf},
677: the convergence 
678: is eventually lost in some simulations; these simulations are no longer
679: solving Einstein's equations.\\
680: 
681: 
682: \begin{figure}
683: \begin{center}
684: \includegraphics[width=5.0in, angle=0]{./figure4}
685: \end{center}
686: \caption{The log of the rms norm for the Hamiltonian constraint
687: in simulations evolving a nonrotating Kerr-Schild black
688: hole with a spatial
689: domain of [$-10M . . . 10M$] at three resolutions:
690: $M/5, M/7.5,$ and $M/10.$ The excision
691: radius was $0.5M$ in all constrained and unconstrained
692: cases. All simulations
693: used a densitized lapse with $p = \frac{1}{3}$.
694: The constraints were
695: solved in the constrained
696: evolution cases every $0.05M$ everywhere on the domain
697: except those points where
698: $r < 2.0M$. Independent residual evaluations for the
699: constrained cases with
700: resolution M/7.5 and M/10 are found in Figure \ref{fig17.pdf}}
701: \label{fig15.pdf}
702: \end{figure}
703: 
704: \begin{figure}
705: \begin{center}
706: \includegraphics[width=5.0in, angle=0]{./figure5}
707: \end{center}
708: \caption{The rms norm of the diagonal
709: spatial components of the Einstein tensor for the
710: M/7.5 and M/10 constrained simulations presented in
711: Figure \ref{fig15.pdf}}
712: \label{fig17.pdf}
713: \end{figure}
714: 
715: \begin{figure}
716: \begin{center}
717: \includegraphics[width=5.0in, angle=0]{./figure6}
718: \end{center}
719: \caption{The z = 0 plane of the
720: Einstein tensor component $G_{xx}$ at time
721: $100M$ for the $M/7.5$ constrained evolution
722: presented in Figures \ref{fig15.pdf}, \ref{fig17.pdf}.
723: }
724: \label{fig20.pdf}
725: \end{figure}
726: 
727: 
728: It is of interest to ask why our approach has not been
729: implemented
730: previously.  A number of factors were at work. It has been
731: universally
732: assumed that the computational overhead of elliptic solvers is
733: excessive. Choptuik \cite{ChoptuikPC} indicates a time
734: penalty of $\times 2$, incurred by a fully constrained $2d$
735: evolution,
736: compared to free evolution. This cost is justified by the
737: much longer
738: physical lifetimes achieved in the ($2d$) evolutions of
739: Ref  \cite{gr-qc0301006}; Choptuik's factor of two in time
740: is considered a small
741: penalty. However, in previous implementations of
742: constrained ($2d$)
743: evolution, one additionally had the problem that the
744: solvers were 
745: restricted, for instance to conformally flat situations.
746: This required 
747: strong gauge constraints on the evolution, and meant that
748: generally one re-solved a strongly nonlinear equation
749: (comparable to the initial value problem), on each time step.
750: But we have found, even working with
751: straightforward package solvers, that the penalty for our
752: approach is only of order 30\%.
753: This is because our constraint solver (developed for Kerr-Schild
754: superposed initial data), is in fact completely general with no form
755: restrictions on the background. (There are, of course, general conditions
756: on 
757: the elliptic equations to allow their solution \cite{restrictions},
758: but we 
759: have encountered no difficulties in working with physically
760: realistic 
761: configurations). 
762: 
763: Thus our elliptic solver can use backgrounds
764: that are
765: strongly nonflat. They may be strongly nonflat, but they are
766: already very close to constraint solution.
767: This is so because they
768: arise from evolving one timestep with an accurate time
769: integrator; we use a package integrator which is typically
770: fourth order
771: accurate in time, and the spatial discretization
772: is also fourth order. There is thus only a very small
773: correction arising from the elliptic solve, and
774: minimal outer iteration is required. Further,
775: we take note of the fact that our explicit time
776: integration has a certain inherent order of truncation error.
777: Thus we do not in fact set the residual
778: limit for the elliptic solve anywhere near machine accuracy.
779: Instead we set it so that it produces errors which are consistent
780: in size with the
781: evolution truncation error.  This reduces the {\it internal}
782: iteration in the solver to a very small number.
783: Finally, computational resources are now becoming
784: adequate for constrained 3-d evolution.
785: $1$Tflop/sec Computers are now accessible,
786: making this work plausible. It
787: will always be the case that constrained evolution is more computation-
788: and memory- intensive than unconstrained, but the time has arrived that
789: interesting constrained evolutions are possible. Further, ongoing
790: computational infrastructure improvements will make this level of
791: computation generally accessible. Well within a decade, desktop access
792: to $10$ Tflop/sec will mean that constrained evolutions at LISA- or
793: LIGO- relevant resolutions will become unexceptional.
794: 
795: 
796: 
797: \section{Discussion}
798: 
799: 
800: 
801: We have demonstrated that the analytical formulation is not critical to long
802: term evolution of single black holes, if the correct constraint subtraction
803: is used.
804: Thus, ``modern" approaches that pose explicitly hyperbolic approaches are
805: not
806: essential; ``traditional" $\dot g$ - $\dot K$ methods produce comparably
807: long
808: evolutions. The evidence seems to be that there are many formulations and
809: subtraction schemes that lead to long-term single black hole stability; we
810: have found an especially simple one.
811: 
812: However, we have found, consistent with theoretical estimates, that the
813: precise subtraction is critical (a fraction to two or three decimals). To
814: attack this problem, we have carried out periodic solution of the elliptic constraint
815: equations as part of
816: the time integration, to enable fully constrained evolutions of
817: the Einstein equations. Our initial results demonstrate dramatic
818: improvement of long
819: term stability of a nonspinning black hole simulation. Because we solve the
820: constraints, constraint subtraction is irrelevant in this case. We are
821: beginning exploration of the constrained evolution
822: approach in spacetimes
823: involving single moving, and multiple interacting black holes. We find
824: substantial improvement from constraint solving in every simulation,
825: but we have not achieved infinite-lived $\dot g - \dot K$ simulations,
826: even with densitized lapse (which is known to improve the hyperbolicity
827: of the system of equations). We have begun an approach where
828: the inner excision is made at a constant coordinate surface that
829: coincides with the apparent horizon (essentially {\it spherical}, or
830: {\it spheroidal} coordinates) near the excision
831: region\cite{mattDiss}. This approach addresses concerns about the validity of the
832: stair-step (LEGO) excision region (\cite{gr-qc/0302072}).
833: (Additionally the best long-lived isolated black hole simulation to
834: date has been carried out by Ref \cite{schKidteuk} with a
835: pseudo-spectral method with apparent horizon conforming coordinates.)
836: 
837: Our approach uses coordinates which are spherical near the hole, with a
838: cartesian 
839: region further away. It is hoped that this approach may exhibit some of the
840: good 
841: properties and long lifetime of the codes described in \cite{schKidteuk}.
842: 
843: 
844: 
845: 
846: 
847: \section*{Acknowledgments}
848: Computations were performed at the Texas Advanced Computing Center
849: at the
850: University of Texas.  This work was supported by NSF grants PHY
851: 0102204 and  PHY~0354842. Additionally, portions of this
852: work were conducted
853: at the Kavli Institute for Theoretical Physics,
854: The University of California at
855: Santa Barbara, under NSF grant PHY99 07947, and at the Laboratory for
856: High Energy Astrophysics, NASA/Goddard Space flight Center,
857: Greenbelt Maryland, with support from the University Space Research
858: Association.  
859: M.~Anderson acknowledges support from
860: a Department of Energy
861: Computational Science Graduate Fellowship administered by the Krell
862: Institute.
863: 
864: %This work is
865: %supported in part by NSF grants PHY~9800722, PHY~9800725, and
866: %PHY~0102204. Computations were performed at the NSF supercomputer
867: %center NCSA and the University of Texas AHPCC.
868: 
869: 
870: \newpage
871: 
872: \begin{thebibliography}{99}
873: 
874: 
875: 
876: 
877: \bibitem{schKidteuk} Mark A. Scheel, Lawrence E. Kidder, Lee Lindblom,
878: Harald P. Pfeiffer, Saul A. Teukolsky, ``Toward stable 3D numerical
879: evolutions of black-hole spacetimes," {\it Phys. Rev.} {\bf D66} 124005
880: (2002)[arXiv:gr-qc/0209115].
881: 
882: 
883: \bibitem{yo}Hwei-Jang Yo, Thomas W. Baumgarte, Stuart L. Shapiro, ``Improved
884: numerical stability of stationary black hole evolution calculations," {\it
885: Phys. Rev.} {\bf D66} 084026 (2002) [arXiv:gr-qc/0209066].
886: 
887: 
888: 
889: 
890: 
891: \bibitem{ADM} R.~Arnowitt, S.~Deser, and C.~Misner in Witten,
892: {\it Gravitation,
893: an Introduction to Current Research} (Wiley, New York 1962).
894: 
895: \bibitem{49Frittelli}Frittelli, S., and Reula, O., ``First-order
896: symmetric-hyperbolic Einstein equations with arbitrary fixed gauge'', {\it
897: Phys. Rev. Lett.}, {\bf 76}, 4667-4670, (1996).
898: 
899: 
900: %\bibitem{Matzner:1999pt}
901: %R.~Matzner, M.~F.~Huq and D.~Shoemaker,
902: %{\it Phys.\ Rev.}  {\bf D59}, 024015 (1999)[arXiv:gr-qc/9805023].
903: 
904: 
905: 
906: \bibitem{KerrSchild} R.~Kerr and A.~Schild, ``Some Algebraically Degenerate
907: Solutions of Einstein's Gravitational Field Equations,'' in
908: {\it Applications of Nonlinear Partial Differential Equations in
909: Mathematical Physics}, Proc. of Symposia B Applied Math., Vol XVII (1965);
910: ``A New Class of Solutions of the Einstein Field Equations'',
911: {\it Atti del Congresso Sulla Relitivita Generale: Problemi Dell'Energia
912: E Onde Gravitazionala} G. Barbera, Ed. (1965).
913: 
914: 
915: \bibitem{SUNDIALS} A.~Hindmarsh, R.~Serban and C.~Woodward, ``SUNDIALS home
916: page",
917: \textit{http://www.llnl.gov/CASC/sundials/} (2002).
918: 
919: \bibitem{CVODES} A.~Hindmarsh and R. Serban, ``User Documentation for
920: CVODES",
921: (Center for Applied Scientific Computer, 
922: Lawrence Livermore National Laboratory, 2002).
923: 
924: \bibitem{KINSOL} A.~Taylor and A. Hindmarsh, ``User Documentation for
925: {KINSOL}, A Nonlinear Solver for Sequentail and Parallel Computers",
926: (Center for Applied Scientific Computer, 
927: Lawrence Livermore National Laboratory, 1998).
928: 
929: \bibitem{gr-qc/0402123}Gabriel Nagy, Omar E. Ortiz and Oscar A. Reula
930: ``Strongly hyperbolic second order Einstein's evolution equations",
931: gr-qc/0402123 (2004).
932: 
933: \bibitem{YP} J.~York and  T.~\ Piran
934:         ``The Initial Value Problem and Beyond'',
935:         \textit{Spacetime and Geometry: The Alfred Schild
936:         Lectures}, R.~Matzner and L.~Shepley Eds.
937:         University of Texas Press, Austin, Texas. (1982);
938:         G.~Cook, ``Initial Data for the Two-Body Problem
939:         of General Relativity'', Ph.D. Dissertation, The University
940:         of North Carolina at Chapel Hill (1990).
941: 
942: 
943: \bibitem{MY} N.~{\'O}.~Murchadha and J.~W.~York, Jr.,
944:                    {\it Phys.\ Rev.}  {\bf D10}, 428 (1974);
945:              N.~{\'O}.~Murchadha and J.~W.~York, Jr.,
946:                    {\it Phys.\ Rev.} {\bf D10}, 437 (1974);
947:              N.~{\'O}.~Murchadha and J.~W.~York, Jr.,
948:                    {\it Gen.\ Relativ.\ Gravit.} {\bf 7} 257 (1976).
949: 
950: \bibitem{York} J.~W.~York, Jr., {\it Phys.\ Rev.\ Lett.} {\bf 82},
951:         1350 (1999).
952:         
953: 
954: 
955: \bibitem{Mathews} J.~R.~Wilson and G.~J.~Mathews,
956:     {\it Phys.~Rev.~Lett.} {\bf 75}, 4161 (1995);
957:     J.~R.~Wilson and G.~J.~Mathews, and P.~Marronetti,
958:     {\it Phys.~Rev.} {\bf D54}, 1317 (1996)
959: 
960: 
961: 
962: 
963: 
964: \bibitem{Bowen+York} J.~Bowen and J.~W.~York,
965:                      {\it Phys. Rev.} {\bf D21}, 2047 (1980).
966: 
967: \bibitem{gr-qc/0002076}
968: Mijan F. Huq, Matthew W. Choptuik and Richard A. Matzner,
969: ``Locating Boosted Kerr and Schwarzschild Apparent Horizons"
970: {\it Physical Review} {\bf D66},~084024~(2002).
971: 
972: \bibitem{ChoptuikPC} M.~Choptuik, Personal communication.
973: 
974: \bibitem{PETSC_1} S.~Balay, K.~Buschelman, W.~D.~Gropp, D.~Kaushik,
975: M.~Knepley, L.~C.~McInnes, B.~F.~Smith and H.~Zhang, ``PETSc home page",
976: \textit{http://www.mcs.anl.gov/petsc} (2001).
977: 
978: \bibitem{PETSC_2} S.~Balay, K.~Buschelman, W.~D.~Gropp, D.~Kaushik,
979: M.~Knepley, L.~C.~McInnes, B.~F.~Smith and H.~Zhang, ``PETSc Users Manual",
980: \textit{ANL-95/11 - Revision 2.1.5} (Argonne National Laboratoy, 2002).
981: 
982: \bibitem{PETSC_3} S.~Balay, K.~Buschelman, W.~D.~Gropp, D.~Kaushik,
983: M.~Knepley, L.~C.~McInnes, B.~F.~Smith and H.~Zhang, ``Efficient Management
984: of 
985: Parallelism in Object Oriented Numerical Software Libraries",
986: \textit{Modern Software Tools in Scientific Computing}, E.~Arge,
987: A.~M.~Bruaset, and H.~P.~Langtangen, editors,
988: (Birkhauser Press, 1997), pp. 163-202.
989: 
990: \bibitem{gr-qc/0302072} Gioel Calabrese, Luis Lehner, David Neilsen,
991: Jorge Pullin, Oscar Reula, Olivier Sarbach, Manuel Tiglio,
992: ``Novel finite-differencing techniques for numerical relativity:
993: application to black hole excision", {\it Classical and Quantum Gravity}
994: {\bf 20} L245-L252 (2003).
995: 
996: \bibitem{projection}G. Toth ``The ${\bf \nabla \cdot}$ {{\bf B}} $= 0$ Constraint in Shock-Capturing
997: Magnetohydrodynamics Codes"
998: {\it Jour. Comp. Phys.} {\bf 161} 605 (2002).
999: 
1000: \bibitem{choptuikSB} Personal communication (2003).
1001: 
1002: 
1003: \bibitem{gr-qc0301006}
1004: M. W. Choptuik,  E. W. Hirschmann,  S. L. Liebling,  and F. Pretorius,
1005: ``An Axisymmetric Gravitational Collapse Code"
1006:  {\it Class. Quant. Grav.} {\bf 20}, 1857-1878 (2003).
1007: 
1008: \bibitem{restrictions} Y. Choquet-Bruhat, J. Isenberg and J. W. York,
1009: {\it Phys. Rev.} {\bf D61} 084034 2000
1010: 
1011: \bibitem{mattDiss}Matthew Anderson, ``Constrained Evolution in Numerical Relativity"
1012: Ph.D. dissertation, The University of Texas at Austin (2004).
1013: 
1014: \end{thebibliography}
1015: \end{document}
1016: \end
1017: