1:
2: \documentclass[twocolumn,showpacs]{revtex4}
3: %\documentstyle[aps,twocolumn,eqsecnum]{revtex}
4:
5: \usepackage{amsmath,epsfig,multirow,delarray}
6: \begin{document}
7:
8: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
9: %%% My Defs:
10: % \newcommand{\ftimes}{f\times} % <-- use Atten. factor f
11: \newcommand{\ftimes}{} % <-- get rid of Atten. factor f
12:
13:
14: % \setlength{\topmargin}{0.0cm} % <--shift down on RHAT 6.2 latex
15: \preprint{CGPG-03/8-1}
16:
17: \title{Numerical stability of the AA evolution system \\ compared to
18: the ADM and BSSN systems}
19:
20: \author{Nina Jansen, Bernd Br\"ugmann, Wolfgang Tichy}
21: \affiliation{
22: Center for Gravitational Physics and Geometry and
23: Center for Gravitational Wave Physics\\
24: Penn State University, University Park, PA 16802
25: }
26:
27: \date{October 20, 2003}
28:
29: \begin{abstract}
30: We explore the numerical stability properties of an evolution system
31: suggested by Alekseenko and Arnold. We examine its behavior on a set
32: of standardized testbeds, and we evolve a single black hole with
33: different gauges. Based on a comparison with two other evolution
34: systems with well-known properties, we discuss some of the strengths and
35: limitations of such simple tests in predicting numerical stability
36: in general.
37: \end{abstract}
38:
39: %\pacs{04.25.Dm, 04.25.Nx, 04.30.Db, 04.70.Bw;}
40: \pacs{
41: 04.25.Dm % Numerical relativity
42: % 04.30.Db, % Wave generation and sources (Gravitational wave theory)
43: % 04.70.Bw, % Classical black holes
44: % 95.30.Sf % Relativity and gravitation (Fundamental aspects of astrophysics)
45: % 97.60.Lf % Black holes (Late stages of stellar evolution)
46: %
47: % Sometimes we want to include preprint numbers, let's put them here
48: \quad Preprint number: CGPG-03/8-1
49: }
50:
51:
52: \maketitle
53:
54:
55:
56:
57: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
58: \section{Introduction}
59:
60: In recent years the quest for finding a numerically stable formulation of
61: the Einstein evolution equations has become more and more intense, see
62: e.g.~\cite{Reula98a,Friedrich:2000qv} for reviews. Effort has been put into
63: finding first order symmetric hyperbolic formulations of the evolution
64: equations, since the properties of such systems can be analyzed
65: mathematically. However, even if an evolution system is symmetric hyperbolic
66: there is no guarantee that its numerical implementation is stable when
67: evolving a highly dynamic black hole spacetime. The
68: Baumgarte-Shapiro-Shibata-Nakamura (BSSN)~\cite{Shibata95,Baumgarte99}
69: system is an example of a system that is not first order symmetric
70: hyperbolic but has nice stability properties. Thus, current mathematical
71: analysis is not sufficient to explore the properties of an evolution system.
72: The system must be implemented and tested numerically before we can draw
73: definite conclusions about its viability for a particular physical
74: application.
75:
76: In \cite{Alekseenko2002}, Alekseenko and Arnold (AA) suggest a first order
77: formulation of the evolution equations that is symmetric hyperbolic when
78: considering only a subset of the variables. In particular, the metric itself
79: and some of its first spatial derivatives are not considered part of the
80: evolution system and are treated as given functions when showing
81: hyperbolicity. It is argued that this is sensible since the metric is
82: derivable from an ordinary differential equation. A distinguishing feature
83: of the AA system is that a minimal number of first derivatives of the metric
84: are introduced as independent variables. The system has only 20 unknowns and
85: no parameters that have to be fixed for hyperbolicity, so it is relatively
86: simple for a symmetric hyperbolic system.
87:
88: We have chosen to implement the AA system numerically and to compare
89: its numerical properties with those of the Arnowitt-Deser-Misner (ADM)
90: \cite{Arnowitt62,York79} and BSSN systems. The ADM system is known to be
91: unstable in many situations, but we include it here since for finite
92: time intervals of evolutions it has been used rather successfully in
93: practice, e.g.\ in 3D black hole
94: simulations~\cite{Anninos94c,Bruegmann97,Brandt00,Alcubierre00b,Baker00b,Baker:2001nu,Baker:2002qf,Baker:2003ds}. The
95: BSSN system can be obtained from a trace-conformal decomposition of
96: the ADM equations, and with suitable techniques it is very stable for
97: black hole evolutions. For example, the first stable evolution for
98: more than $100000M$ of a single Schwarzschild black hole in a
99: (3+1)-dimensional code without adapted coordinates was obtained with a
100: modified BSSN system~\cite{Alcubierre00a}.
101: Both the ADM and BSSN systems are not first order. However, there exist first
102: order versions of the BSSN system which are symmetric
103: hyperbolic~\cite{Alcubierre99c,Frittelli99,Friedrich:2000qv,Sarbach02a}, if the
104: densitized lapse and shift are considered given functions.
105: Straightforward first order forms of ADM are only
106: weakly hyperbolic, which implies certain numerical instabilities
107: (see e.g.~\cite{Calabrese02a}).
108:
109: The second order version of the BSSN system shares with the AA system
110: the property that a subsystem of it is indeed symmetric hyperbolic. A
111: crucial step in the construction of BSSN is the introduction of the
112: contracted Christoffel symbol of the conformal metric,
113: $\tilde\Gamma^i$, as an independent variable. The evolution equation
114: for the extrinsic curvature then contains derivatives of the metric
115: only in the form of a Laplace operator, so that the metric obeys a
116: wave equation if $\tilde\Gamma^i$ is considered a prescribed variable,
117: ignoring its own evolution equation. This partial hyperbolicity of
118: the BSSN system may well be a crucial ingredient in its success as
119: evolution system. Hence the question arises whether symmetric
120: hyperbolicity in a subsystem implies a numerical advantage for the AA
121: system.
122:
123: There is a variety of numerical tests that one can perform on an
124: evolution system. A complex and important issue is that seemingly
125: minor changes in the test conditions and implementation of the system
126: can lead to very different conclusions about
127: stability. In~\cite{Alcubierre2003:mexico-I}, an important step is
128: taken towards creating a set of tests that can serve as a standardized
129: benchmark for stability. Our discussion of the AA system in comparison
130: to ADM and BSSN can also be viewed as a contribution to the ongoing
131: development of such benchmarks. On the one hand, we report on the
132: performance of our particular implementation of ADM and BSSN, where a
133: large body of prior work allows us to judge how representative the
134: current benchmark is. On the other hand, we apply these tests to a
135: new evolution system about which nothing is known so far from numerical
136: experiments, and the question is, for example, whether one can predict the
137: usefulness of the AA system for black hole evolutions. We therefore
138: also discuss results for the evolution of a single black hole that
139: go beyond the current set of benchmarks in~\cite{Alcubierre2003:mexico-I}.
140:
141: The paper is organized as follows. In Sec.~\ref{secAA}, we write out
142: the AA system explicitly, in notation that is more familiar to
143: numerical relativists, and we show that the complete system is not
144: symmetric hyperbolic. We introduce a time-independent conformal
145: rescaling of the AA system that we will need later for black hole
146: evolutions. In Sec.~\ref{secTests}, we perform three tests from the
147: recently proposed test suite \cite{Alcubierre2003:mexico-I} for
148: numerical evolutions, the robust stability test, the gauge stability
149: test, and the linear wave stability test. Finally, we evolve a single
150: black hole space time in Sec.~\ref{secBH}, since this is the physics
151: example that we are most interested in. We conclude with a discussion
152: in Sec.~\ref{secDiscuss}.
153:
154:
155:
156:
157:
158:
159: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
160: \section{The AA system}
161: \label{secAA}
162:
163: In~\cite{Alekseenko2002}, the evolution system is given in compact
164: notation. Here we write it out in the form in which we have
165: implemented it:
166: \begin{widetext}
167: \begin{eqnarray}
168: \label{AAeqn_first}
169: f_{ijk} &=& \frac{1}{2\sqrt{2}} \left(g_{ik,j} - g_{jk,i} + \left(\left(g_{pj,q}-g_{pq,j}\right)g_{ik} - \left(g_{pi,q} - g_{pq,i}\right)g_{jk}\right) g^{pq} \right) ,\\
170: w_{ij} &=& \frac{1}{2}\left(\beta^{p}_{\;,j} g_{pi} + \beta^{p}_{\;,i}
171: g_{pj}\right) ,\\
172: K &=& g^{ij} K_{ij} ,\\
173: c^{1)}_{ij} &=& \frac{1}{4} \left(2 g_{pq,j} g^{pq}_{\;\;,i} - g_{pq,i} g^{pq}_{\;\;,j} + 2 g_{pj,i} g^{pq}_{\;\;,q} - 2 g_{pj,q}
174: \left(g^{pq}_{\;\;,i} + g_{ri,s} \left(g^{ps} g^{qr} - g^{pr}
175: g^{qs}\right)\right) - \right.\\ \nonumber
176: &\quad& \left. g_{ij,s} g_{pq,r} g^{pq} g^{rs} + g_{pq,r} g_{si,j}
177: g^{pq} g^{rs} + g_{pq,r} g_{sj,i} g^{pq} g^{rs}\right) ,\\
178: c^{2)}_{ij} &=& \frac{1}{4}\left(2 c^{1)}_{ij} + 2 c^{1)}_{ji} - g_{pj,q} g^{rs}_{\;\;,s} g_{ri} g^{pq} +
179: g_{pq,j} g^{rs}_{\;\;,s} g_{ri} g^{pq} - g_{pi,q} g^{rs}_{\;\;,s}
180: g_{rj} g^{pq} + g_{pq,i} g^{rs}_{\;\;,s} g_{rj} g^{pq} + \right. \\
181: \nonumber &\quad& \left.
182: g_{pq,r} g^{rs}_{\;\;,j} g_{si} g^{pq} + g_{pq,r} g^{rs}_{\;\;,i}
183: g_{sj} g^{pq} - g_{pq,r} g^{qs}_{\;\;,j} g_{si} g^{pr} -
184: g_{pq,r} g^{qs}_{\;\;,i} g_{sj} g^{pr} - g_{pj,q} g^{qr}_{\;\;,s}
185: g_{ri} g^{ps} - \right. \\
186: \nonumber &\quad& \left.
187: g_{pq,j} g^{qr}_{\;\;,s} g_{ri} g^{ps} -
188: g_{pi,q} g^{qr}_{\;\;,s} g_{rj} g^{ps} - g_{pq,i} g^{qr}_{\;\;,s}
189: g_{rj} g^{ps} + 2 g_{pj,q} g^{pr}_{\;\;,s} g_{ri} g^{qs} +
190: 2 g_{pi,q} g^{pr}_{\;\;,s} g_{rj} g^{qs} - \right. \\
191: \nonumber &\quad& \left.
192: 4 g_{ij,s} g_{pq,r} g^{pr} g^{qs} +
193: 4 g_{ij,s} g_{pq,r} g^{pq} g^{rs}\right) ,\\
194: c^{4)}_{ij} &=& \frac{1}{2}\left(2\alpha c^{2)}_{ij} - 2 \alpha_{,i,j} - \alpha_{,p} g_{ij,q} g^{pq} +
195: \alpha_{,p} g_{qi,j} g^{pq} + \alpha_{,p} g_{qj,i} g^{pq} +
196: 2\alpha K K_{ij} + \right. \\
197: \nonumber &\quad& \left.
198: 2 \beta^{p}_{\;,j} K_{pi} + 2 \beta^{p}_{\;,i} K_{pj} -
199: 4\alpha g^{pq} K_{pi} K_{qj}\right) ,\\
200: c^{5)} &=& -\frac{1}{4} \left(g_{pq,r} \left(2 g^{qr}_{\;\;,s} g^{ps} - g^{pq}_{\;\;,s} g^{rs} + g_{st,u} g^{pq} g^{ru} g^{st}\right)\right) \\
201: B_{ij} &=& \frac{1}{2} \left(2 c^{4)}_{ij} -\alpha \left(g_{ij,s} g_{pq,r} \left(-\left(g^{ps} g^{qr}\right) +
202: g^{pq} g^{rs}\right) + g_{ij} \left(c^{5)} + K^2 - K_{pq}
203: K^{pq}\right)\right)\right) , \\
204: c^{6)}_{ijk} &=& \frac{1}{2 \sqrt{2}} \left(\beta^{p}_{\;,q} \left(g_{rj,p} g_{ik} - g_{ri,p} g_{jk}\right) g^{qr} +
205: \beta^{p}_{\;,j} \left(g_{ik,p} - g_{qr,p} g_{ik}
206: g^{qr}\right) + \beta^{p}_{\;,i} \left(-g_{jk,p} + g_{qr,p} g_{jk}
207: g^{qr}\right)\right) ,\\
208: c^{7)}_{ijk} &=& \frac{1}{2}
209: \left(2 c^{6)}_{ijk} + \sqrt{2} \left(w_{ik,j} - w_{jk,i} +
210: w_{pj,q} g_{ik} g^{pq} - w_{pq,j} g_{ik} g^{pq} -
211: w_{pi,q} g_{jk} g^{pq} + w_{pq,i} g_{jk} g^{pq} - \right. \right. \\
212: \nonumber &\quad& \left. \left.
213: \alpha \left( g_{pj,q} g^{pq} K_{ik} + g_{pq,j} g^{pq} K_{ik} +
214: g_{pi,q} g^{pq} K_{jk} - g_{pq,i} g^{pq} K_{jk} +
215: g_{pj,q} g_{ik} K^{pq} - g_{pq,j} g_{ik} K^{pq} - \right. \right. \right.\\
216: \nonumber &\quad& \left. \left. \left.
217: g_{pi,q} g_{jk} K^{pq} + g_{pq,i} g_{jk} K^{pq}\right) +
218: g_{pj,q} g^{pq} w_{ik} - g_{pq,j} g^{pq} w_{ik} -
219: g_{pi,q} g^{pq} w_{jk} + g_{pq,i} g^{pq} w_{jk} - \right. \right.\\
220: \nonumber &\quad& \left. \left.
221: g_{pj,q} g_{ik} w^{pq} + g_{pq,j} g_{ik} w^{pq} +
222: g_{pi,q} g_{jk} w^{pq} - g_{pq,i} g_{jk} w^{pq}\right)\right) ,\\
223: C_{ijk} &=& \frac{1}{4} \left(4 c^{7)}_{ijk} + \sqrt{2} \left(2 K \alpha_{,j} g_{ik} - 2 K \alpha_{,i} g_{jk} -
224: 2\alpha g^{pq}_{\;\;,q} g_{jk} K_{pi} + 2\alpha g^{pq}_{\;\;,q} g_{ik}
225: K_{pj} -\alpha g^{pq}_{\;\;,j} g_{ik} K_{pq} + \right. \right.\\
226: \nonumber &\quad& \left. \left.
227: \alpha g^{pq}_{\;\;,i} g_{jk} K_{pq} + 2 \alpha_{,p} g_{jk} g^{pq}
228: K_{qi} - 2 \alpha_{,p} g_{ik} g^{pq} K_{qj} -
229: \alpha g_{pq,r} g_{jk} g^{pq} g^{rs} K_{si} +
230: \alpha g_{pq,r} g_{ik} g^{pq} g^{rs} K_{sj}\right)\right) ,\\
231: \partial_0 g_{ij} &=& - 2 \alpha K_{ij} + 2 w_{ij} ,\\
232: \partial_0 K_{ij} &=& \frac{1}{2} \left(2 B_{ij} + \sqrt{2}\alpha \left(g^{pq}_{\;\;,q} \left(f_{pij} + f_{pji}\right) +
233: f_{pij,q} g^{pq} + f_{pji,q} g^{pq} -
234: g^{pq}_{\;\;,r} f_{psj} g_{qi} g^{rs} + g^{pq}_{\;\;,r} f_{sjp}
235: g_{qi} g^{rs} - \right. \right.\\
236: \nonumber &\quad& \left. \left.
237: g^{pq}_{\;\;,r} f_{psi} g_{qj} g^{rs} +
238: g^{pq}_{\;\;,r} f_{sip} g_{qj} g^{rs}\right)\right) ,
239: \\
240: \partial_0 f_{ijk} &=& C_{ijk} + \frac{1}{\sqrt{2}} \left(-\alpha
241: K_{ik,j} +\alpha K_{jk,i} - \alpha_{,j} K_{ik} + \alpha_{,i}
242: K_{jk}\right) .
243: \label{AAeqn_last}
244: \end{eqnarray}
245: \end{widetext}
246: The operator $\partial_0 = \partial_t - \beta^p\partial_p$ is defined
247: in terms of ordinary partial derivatives for all variables.
248:
249: The new variable $f_{ijk}$ is anti-symmetric in $i$ and $j$ and also
250: satisfies the cyclic property $f_{ijk} + f_{jki} + f_{kij} = 0$. This
251: means that $f_{ijk}$ represents 8 independent variables. The total
252: number of 20 variables is thus smaller than in the case of most first
253: order formulations (which usually involve 30 or more variables), but
254: larger than for the BSSN system, which has 15 independent variables.
255:
256: In addition, the equations in the AA system are more complex than in
257: the BSSN system, so that the AA evolution system takes roughly twice
258: as long to run the same number of iterations, even though we have made
259: an effort to implement the AA system in a numerically optimal way. The
260: simulations were carried out with the BAM code, which is a rewritten
261: version of the code used in~\cite{Bruegmann97}. Part of
262: BAM is a Mathematica script to convert tensor equations to C code, and
263: the AA equations were generated this way using Mathematica and
264: MathTensor.
265:
266:
267:
268: \subsection{Hyperbolicity of the AA system}
269: \label{sec:hyper}
270: Looking at Eqs.\ (\ref{AAeqn_first})-(\ref{AAeqn_last})
271: and assuming that $\alpha$ and $\beta^i$ are prescribed given functions,
272: we see that the system has the form
273: \begin{equation}
274: \label{AA_form}
275: \partial_t u + A^i(u) u_{,i} + v(u)
276: + w\left(m^{ij}(u) u_{,i}^T Q(u) u_{,j} \right) = 0 .
277: \end{equation}
278: Here
279: \begin{equation}
280: u=\begin{pmatrix} g \\
281: K \\
282: f \\ \end{pmatrix}
283: \end{equation}
284: is the solution vector with spatial indices suppressed, and
285: \begin{equation}
286: A^i(u)=\begin{pmatrix} 0 & 0 & 0 \\
287: r^i(u) &a^i(u) & c^i(u) \\
288: s^i(u) &c^i(u) & b^i(u) \\ \end{pmatrix}
289: \end{equation}
290: is a matrix which contains a symmetric submatrix
291: \begin{equation}
292: S^i(u)=\begin{pmatrix} a^i(u) & c^i(u) \\
293: c^i(u) & b^i(u) \\ \end{pmatrix} .
294: \end{equation}
295: In addition, Eq.~(\ref{AA_form}) contains the two vector valued
296: functions $v$ and $w$, where the argument of $w$ depends on
297: \begin{equation}
298: u_{,i}=\begin{pmatrix} g_{,i} \\
299: K_{,i} \\
300: f_{,i} \\ \end{pmatrix}
301: \end{equation}
302: and its transpose $u_{,i}^T$,
303: and also on another matrix $Q(u)$,
304: which is of the simple form
305: \begin{eqnarray}
306: Q(u)=\begin{pmatrix} q(u) & 0 & 0 \\
307: 0 & 0 & 0 \\
308: 0 & 0 & 0 \\ \end{pmatrix} ,
309: \end{eqnarray}
310: so that the argument of $w$ in fact only depends on squares of first
311: derivatives of $g_{ij}$.
312:
313: From Eq.~(\ref{AA_form}) it is immediately apparent that the system is
314: of first order form, as no second derivatives of $u$ appear.
315: Nevertheless the system differs from the standard first order form by
316: the term $w$. This means that the standard theorems about
317: well-posedness (see e.g.~\cite{Reula98a})
318: are not applicable, and that we cannot easily compute
319: characteristic speeds and modes of the system. However, if we
320: linearize the system around any background $u^B$, all terms of the
321: form $w$ drop out. This can be seen as follows. Assume that
322: \begin{equation}
323: \label{linearAnsatz}
324: u=u^B + u^L = \begin{pmatrix} g^B +g^L \\
325: K^B +K^L \\
326: f^B +f^L \\ \end{pmatrix} ,
327: \end{equation}
328: where $u^L$ is a small perturbation to the background $u^B$.
329: Then
330: \begin{eqnarray}
331: && w\left(m^{ij}(u) u_{,i}^T Q(u) u_{,j} \right) \nonumber \\
332: &&= w\left(m^{ij}(u) g_{,i} q(u) g_{,j} \right) \nonumber \\
333: &&= w\left(m^{ij}(u) g^B_{,i} q(u) g^B_{,j} \right) + \nonumber \\
334: &&\quad\;\; w\left(m^{ij}(u) ( g^B_{,i} q(u) g^L_{,j} +
335: g^L_{,i}q(u) g^B_{,j}) \right) + O\left( (g^L)^2 \right) \nonumber \\
336: &&= w\left(m^{ij}(u) g^B_{,i} q(u) g^B_{,j} \right) +
337: d^i(u^B) g^L_{,i} +O\left( (g^L)^2 \right),
338: \end{eqnarray}
339: and hence Eq.~(\ref{AA_form}) becomes
340: \begin{equation}
341: \label{AA_form_linear}
342: \partial_t u^L + \bar{A}^i u^L_{,i} + \bar{v}(u^L)
343: + O\left( (u^L)^2 \right) + \bar{w}(u^B) = 0 ,
344: \end{equation}
345: which is now of standard first order form, but with a modified
346: matrix
347: \begin{equation}
348: \label{A_linear}
349: \bar{A}^i=\begin{pmatrix} 0 & 0 & 0 \\
350: \bar{r}^i &a^i(u^B) & c^i(u^B) \\
351: \bar{s}^i &c^i(u^B) & b^i(u^B) \\ \end{pmatrix}
352: .
353: \end{equation}
354: Looking at Eq.~(\ref{AA_form_linear}) and (\ref{A_linear}), it is
355: immediately clear that the system is not symmetric hyperbolic. In
356: fact, since the matrix $S^i(u^B)$ has several zero eigenvalues,
357: the full system in general is unlikely to have a complete set of
358: eigenvectors and thus to be strongly hyperbolic.
359:
360: The exception is linearizion around flat space where
361: $\bar{r}^i$ and $\bar{s}^i$
362: in Eq.~(\ref{A_linear}) vanish, so that the system
363: (\ref{AA_form_linear}) becomes symmetric hyperbolic in this special
364: case. Also, since $S^i(u)$ is symmetric, the subsystem of $K_{ij}$
365: and $f_{ijk}$ with $g_{ij}$ considered as a prescribed variable is
366: symmetric hyperbolic, which holds true even for
367: the non-linearized system (\ref{AA_form}).
368: Alekseenko and Arnold \cite{Alekseenko2002} emphasize this last point and
369: stress that the evolution equation for $g_{ij}$ is (as usual) just an
370: ordinary differential equation. Yet in numerical evolutions the latter
371: property may be unimportant, since
372: $g_{ij}$ has to be evolved along with the other variables and cannot be
373: prescribed. In addition, it is not clear if the fact that the evolution
374: system (\ref{AA_form}) contains a symmetric hyperbolic subsystem has any
375: bearing on the stability properties of the full system.
376:
377:
378:
379:
380: \subsection{Conformal version of the AA system}
381:
382: In Sec.~\ref{secBH}, we report on numerical evolutions of black hole
383: puncture data \cite{Brandt97b,Bruegmann97,Alcubierre02a}.
384: In order to numerically evolve puncture data without excision the AA
385: system has to be modified. The reason is that e.g.\ for two punctures
386: the metric has the form
387: \begin{equation}
388: \label{punc_metric}
389: g_{ij} = \phi^{4} \bar{g}_{ij}
390: \end{equation}
391: with conformal factor
392: \begin{equation}
393: \phi = 1 + \frac{m_1}{2r_1} + \frac{m_2}{2r_2} + u ,
394: \end{equation}
395: where $r_1$ and $r_2$ are distances from the punctures, $m_1$ and
396: $m_2$ are the bare puncture masses, and $u$ is finite. From
397: Eq.~(\ref{punc_metric}) it is apparent that the physical metric
398: diverges like
399: \begin{equation}
400: g_{ij} \sim \frac{1}{r^4}
401: \end{equation}
402: at each puncture, so that finite differencing calculations across or
403: near any puncture will fail. The same problem also occurs in the
404: variable
405: \begin{equation}
406: f_{ijk} \sim (g_{ik,j} - g_{jk,i}) \sim \frac{1}{r^5} .
407: \end{equation}
408: For this reason we do not evolve the variables $g_{ij}$ and $f_{ijk}$
409: directly, rather we rescale $g_{ij}$ and $f_{ijk}$ by the time
410: independent conformal factor
411: \begin{equation}
412: \psi = 1 + \frac{m_1}{2r_1} + \frac{m_2}{2r_2}
413: \end{equation}
414: such that the rescaled quantities
415: \begin{eqnarray}
416: \label{scaled_g}
417: \bar{g}_{ij} &=& \psi^{-4} g_{ij}, \\
418: \label{scaled_f}
419: \bar{f}_{ijk} &=& \psi^{-6} f_{ijk}
420: \end{eqnarray}
421: become finite at the puncture.
422: Furthermore, we also introduce the rescaled extrinsic curvature
423: \begin{equation}
424: \label{scaled_K}
425: \bar{K}_{ij} = \psi^{-4} K_{ij} ,
426: \end{equation}
427: in order to remove divergences in $K_{ij}$, which in the case of
428: punctures with spin behaves as
429: \begin{equation}
430: {K}_{ij} \sim \frac{1}{r} .
431: \end{equation}
432:
433: Since $\psi$ is an a priori prescribed time independent function,
434: the principal part of the system of evolution equations
435: (\ref{AAeqn_first})-(\ref{AAeqn_last}) remains unchanged if we use the
436: rescalings in Eqs.~(\ref{scaled_g}), (\ref{scaled_f}), and
437: (\ref{scaled_K}) to express the system in terms of the new variables
438: $\bar{g}_{ij}$, $\bar{K}_{ij}$, and $\bar{f}_{ijk}$. In addition, all
439: terms involving $\bar{g}_{ij}$, $\bar{K}_{ij}$, $\bar{f}_{ijk}$, and
440: their derivatives are finite in the new system so that finite
441: differencing can be used without trouble. There are, however,
442: additional terms containing spatial derivatives of $\psi$, which
443: cannot be computed using finite differencing. Yet, since $\psi$ is a
444: known function we can use analytic expressions for its
445: derivatives. Furthermore, all such derivative terms also contain
446: appropriate powers of $\psi$ which make them finite.
447:
448:
449:
450:
451:
452:
453: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
454: \section{Stability tests}
455: \label{secTests}
456:
457: We have used some of the stability testbeds suggested in
458: \cite{Alcubierre2003:mexico-I}, also known as the ``apples with apples'' tests,
459: to explore the properties of the AA system. We also show test results
460: for the ADM and BSSN systems for comparison.
461:
462:
463:
464:
465: \subsection{Robust stability test}
466:
467: The purpose of the robust stability test is to determine how an
468: evolution system will handle random errors that are bound to occur at
469: machine-precision level. Random constraint violating initial data in
470: the linearized regime is used to simulate this machine error.
471:
472: All apples with apples test are run on a full-3D grid, but in this
473: case with periodic boundary conditions in all directions. The 3D
474: domain is a 3-torus, and here we only use 3 grid-points in the $y$- and
475: $z$-directions. The parameters are:
476: \begin{itemize}
477: \item Simulation domain: $x \in [-0.5, +0.5]$
478: \item Number of grid points in each direction:
479: $n_x=50\rho$, $n_y=n_z=3$, with $\rho=1,2,4$
480: \item Courant factor $= 0.5$
481: \item Iterations: $100000\rho$, output every $100\rho$ iterations
482: (corresponding to 1 crossing time)
483: \item Gauge: harmonic, i.e.
484: $\partial_t \alpha = -\alpha^2 \mbox{tr}K$, $\beta^i=0$
485: \end{itemize}
486:
487: The initial data is given by
488: \begin{eqnarray}
489: g_{ij} &=& \delta_{ij} + \varepsilon_{ij} , \\
490: K_{ij} &=& 0 , \\
491: \alpha &=& 1 , \\
492: \beta^i &=& 0 ,
493: \end{eqnarray}
494: where $\varepsilon_{ij}$ is a random number with a probability distribution,
495: which is uniform in the interval $(-10^{-10}/\rho^2,+10^{-10}/\rho^2)$.
496: After a small number of timesteps, the random noise in $g_{ij}$ will have
497: propagated into all other evolved quantities, except for $\beta^i$, which
498: will remain identically zero for all time. Hence our initial data differ
499: slightly form the ones in \cite{Alcubierre2003:mexico-I},
500: who add random noise to all quantities which need initialization.
501:
502: There is one obvious problem with the robust stability test. The
503: initial data are random, and we must carefully check that the random
504: number generator we use does not introduce errors because the numbers
505: are not totally random. We use C's ``random'' function on Linux Redhat 7.3,
506: a pseudo-random number generator based on a non-linear additive
507: feedback algorithm that avoids the short-comings of some of the older
508: implementations of the ``rand'' function.
509: As a seed for the random number generator we use the time where the
510: subroutine is called plus the clock cycle of the processor on which it
511: is called. Since our code is parallelized and each processor uses its
512: own seed, the actual random numbers on the grid depend on the number
513: of processors. This could be avoided by additional coding, however,
514: the qualitative result of the robust stability test is not expected to
515: depend on the actual random numbers. We have run the test several
516: times on different numbers of processors. We have also tried increasing
517: the size of the domain, both in the x-direction and the y- and
518: z-directions (which results in a different number of random numbers being
519: generated). We find that although this does change the results, many
520: features are robust and do not change with these variations. We will
521: only comment on these robust features of the test.
522:
523: ADM and BSSN results are similar to those reported in
524: \cite{Alcubierre2003:mexico-I}. ADM grows exponentially
525: (Fig.~\ref{fig_admrobust}), while BSSN is stable (Fig.~\ref{fig_bssnrobust}).
526: Note that ADM clearly shows the signature of a grid mode, since when
527: plotting versus the number of iterations almost identical exponential
528: growth is obtained for the three different resolutions.
529:
530: In~\cite{Bona:2003qn}, the ADM system is run with a Courant factor
531: of only $0.03$ with the result that it is stable for much longer, up
532: to 200 crossing times.
533: We have lowered the Courant factor in the ADM
534: run to $0.25$ and found that we do not see exponential growth in the
535: Hamiltonian constraint. When plotting $\mbox{tr}K$ we see some growth,
536: similar to Fig.~2 in~\cite{Bona:2003qn}, but we do not encounter a blowup
537: (we ran the coarsest resolution to 10000 crossing times, and did not
538: encounter exponential growth). The results are shown in
539: Fig.~\ref{fig_admrobust025}. We also tried with Courant factor $0.03$
540: as in~\cite{Bona:2003qn}, and this did not show any exponential blowup (the
541: run was stopped at 600 crossing times). Preliminary experiments indicate
542: that the transition between stable and unstable Courant factors
543: for ADM lies between $0.4$ and $0.5$. This deserves further investigation.
544:
545: The results for AA are shown in Fig.~\ref{fig_AArobust}. AA is
546: qualitatively stable for the 1000 crossing times suggested for this
547: test, but oscillations in the maximum and minimum of the Hamiltonian
548: occur. There are two types of oscillations, one with a long wavelength
549: and one with a short wavelength. We can see that the short wavelength
550: oscillations dampen out. The long wavelength of the medium and high
551: resolution runs seems to be the same, while the wavelength in the low
552: resolution run is about half of the wavelength in the other runs. As
553: pointed out earlier, these features are robust when changing the
554: domain-size, but the amplitude of the oscillations are affected by
555: this change.
556:
557: Given the presence of these slow oscillations, we ran another set of tests
558: where we evolve for 10000 crossing times.
559: Fig.~\ref{fig_AArobustlong} shows the results. We can see that at late
560: times, the constraint violation for all 3 resolutions are growing, and
561: this suggests that the AA system is stable for random noise initial
562: data for a long time, but it will eventually become unstable. We also
563: ran the test of the BSSN system to 10000 crossing times but found no
564: instabilities or any indication of longterm growing oscillations in
565: this case.
566:
567: We tried lowering the Courant factor for the AA runs. We ran the
568: coarsest resolution to 10000 crossing times with a Courant factor of
569: 0.25 and found that the growth in the Hamiltonian constraint still
570: happens, but it happens later than in the run with the Courant factor
571: of 0.5. Lowering the Courant factor for the AA system does not change
572: the general features of the oscillations.
573:
574:
575:
576:
577: \subsection{Gauge wave stability test}
578:
579: In this test we look at the ability of the evolution systems to
580: handle gauge dynamics. This is done by considering flat Minkowski
581: space in a slicing where the 3-metric $g_{ij}$ is time dependent. The
582: gauge wave is then given by
583: \begin{eqnarray}
584: g_{xx} &=& 1 + A \sin \left( \frac{2 \pi (x - t)}{d} \right), \\
585: g_{yy} = g_{zz} &=& 1, \\
586: g_{xy} = g_{xz} = g_{yz} &=& 0, \\
587: K_{xx} &=& -\frac{\pi A}{d} \frac{ \cos \left( \frac{2 \pi (x - t)}{d}
588: \right) }{ \sqrt{1 + A \sin \left( \frac{2 \pi (x - t)}{d} \right) }
589: },\\
590: K_{ij} &=& 0 \qquad \textrm{ otherwise} \\
591: \alpha &=& \sqrt{1 + A \sin \left( \frac{2 \pi (x - t)}{d} \right)} , \\
592: \beta^i &=& 0.
593: \end{eqnarray}
594: Here $d$ is the size of the domain in the $x$-direction
595: and $A$ is the amplitude of the wave.
596: Since this wave propagates along the $x$-axis and all derivatives are
597: zero in the $y$- and $z$-directions, the problem is essentially one
598: dimensional.
599:
600: These are the parameters we use in our gauge stability tests:
601: \begin{itemize}
602: \item Simulation domain: $x \in [-0.5, +0.5]$
603: \item Points: $n_x=50\rho$, $n_y=n_z=3$, $\rho=1,2,4$
604: \item Courant factor $= 0.25$
605: \item Iterations: $200000\rho$, output every $200\rho$ iterations
606: (corresponding to 1 crossing time)
607: \item $A=0.1$ and $A=0.01$
608: \item Gauge: harmonic, i.e.
609: $\partial_t \alpha = -\alpha^2 \mbox{tr}K$, $\beta^i=0$
610: \end{itemize}
611:
612: Since we know the analytical solution at all times, we can compare our
613: numerical results to the analytical results, see
614: also~\cite{Calabrese02a}. As well as testing if the system has
615: exponentially growing modes, we can check the convergence
616: of the numerical result to the analytical solution with increasing
617: resolution.
618:
619: Fig.~\ref{fig_AAgauge} shows that for a gauge wave amplitude
620: $A=0.1$ the AA system converges as expected for a finite time
621: interval, but there is exponential growth and the run crashes after about
622: 100 crossing times.
623: For $A=0.01$ (Fig.~\ref{fig_AAgaugeA001}) it takes longer
624: for this non-convergence to show, about 1000 crossing times,
625: but it is still there. If the AA
626: system together with the given gauge equations
627: was symmetric hyperbolic, and if one had a stable
628: discretization scheme, then the result would be
629: convergent. Conversely, assuming that ICN is an appropriate
630: discretization scheme, we would conclude that the AA system
631: in harmonic gauge is not symmetric hyperbolic, which agrees with our
632: result in Sec.~\ref{sec:hyper}.
633:
634: The ADM system shows no growth of the constraints at all, and the
635: constraint violation remains at machine precision. This is probably
636: because for a 1-dimensional gauge wave the ADM system simplifies such
637: that no constraints violating modes are possible.
638:
639: The results for the BSSN system are shown in
640: Fig.~\ref{fig_bssngauge}. The BSSN system becomes unstable and
641: non-convergent rather quickly (on the order of 100 crossing
642: times). This is somewhat surprising since the BSSN system has been
643: very successful in single black hole evolutions and neutron star
644: evolutions~\cite{Alcubierre02a,Duez:2002bn}, so we expect it to be
645: able to handle gauge dynamics. However, part of the robustness of BSSN
646: can be attributed to the fact that constraint violating modes have a
647: finite speed~\cite{Alcubierre99e,Knapp02a} and can propagate out of
648: the grid for example for radiative boundary conditions. However, on
649: our grid with periodic boundaries, when a constraint violating mode
650: appears it will stay on the grid, which is probably the reason that
651: the BSSN system fails this test. Thus, this test by itself cannot
652: determine whether a system can handle gauge dynamics, but must be
653: accompanied by other tests without periodic boundaries. In particular,
654: there is no contradiction to the observed stability of BSSN for single
655: black holes with radiative boundaries.
656:
657:
658:
659:
660:
661: \subsection{Linear wave stability test}
662:
663: In this section we investigate the ability of the evolution systems to
664: propagate the amplitude and phase of a traveling gravitational
665: wave. The amplitude of this wave is sufficiently small so that we are
666: in the linear regime. This test reveals effects of numerical
667: dissipation and other sources of inaccuracy in the evolution
668: algorithm.
669:
670: The initial 3-metric and extrinsic curvature are obtained from
671: \begin{equation}
672: ds^2= - dt^2 + dx^2 + (1+b) \, dy^2 + (1-b) \, dz^2,
673: \end{equation}
674: where
675: \begin{equation}
676: b = A \sin \left( \frac{2 \pi (x-t)}{d}\right),
677: \end{equation}
678: $d$ is the size of the propagation domain, and $A$ is the amplitude of
679: the wave. The extrinsic curvature tensor is given by
680: \begin{eqnarray}
681: K_{yy} &=& -\frac{A \pi}{d} \cos\left(\frac{2 \pi x}{d}\right) , \\
682: K_{zz} &=& \frac{A \pi}{d} \cos\left(\frac{2 \pi x}{d}\right) , \\
683: K_{ij} &=& 0, \qquad \text{otherwise}.
684: \end{eqnarray}
685:
686: These are the parameters of our run:
687: \begin{itemize}
688: \item Simulation domain: $x \in [-0.5, +0.5]$
689: \item Points: $n_x=50\rho$, $n_y=n_z=3$, $\rho=1,2,4$
690: \item Courant factor $= 0.25$
691: \item Iterations: $200000\rho$, output every $200\rho$ iterations
692: (corresponding to 1 crossing time)
693: \item $A=10^{-8}$
694: \item Gauge: geodesic, i.e.\ $\alpha = 1, \beta^i = 0$.
695: \end{itemize}
696:
697: Figs.~\ref{fig_linearcomp},\ref{fig_linearnorm}, show the results for
698: the ADM, AA, and BSSN systems, respectively.
699: Fig.~\ref{fig_linearcomp} is a comparison of the numerical wave to the
700: analytical wave at 100 crossing times. We see that the AA numerical
701: wave travels slightly faster than the ADM and BSSN numerical
702: waves. Fig.~\ref{fig_linearnorm} shows the
703: $L_2$-norm of the difference between the numerical and analytical
704: linear waves as a function of time, and again the wave in the AA
705: system travels faster than in the other systems. The
706: Hamiltonian constraint has a value of about $10^{-8}$ at the end of
707: the run for the AA system, so we are still well within the linear
708: regime. It is also worth noting that in the ADM system the constraint
709: violation is constant throughout the evolution, while the constraints
710: grow slightly for the other two systems (but the maximum violation is
711: of the order $10^{-8}$ for the entire run, for all resolutions).
712:
713:
714:
715:
716:
717: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
718: \section{Numerical tests involving a single Schwarzschild black hole}
719: \label{secBH}
720:
721: In this Section we present numerical results for the evolution of a single
722: Schwarzschild black hole in two different gauges.
723:
724:
725: \subsection{Geodesic slicing}
726:
727: The initial data for this test is a Schwarzschild black hole in
728: isotropic coordinates. We use geodesic slicing to evolve the initial
729: data. In geodesic slicing, the coordinate lines correspond to freely
730: falling observers, and the resulting slicing of Schwarzschild can be
731: expressed analytically in terms of Novikov coordinates,
732: e.g.~\cite{Frolov98}, Chapter 2.7.2, and~\cite{Bruegmann96}, which we
733: transform to isotropic coordinates for direct comparison with the
734: numerical results. To this end, we numerically solve the following
735: implicit equation for the Schwarzschild area radial
736: coordinate $R=R(\tau,R_\text{max})$,
737: \begin{eqnarray}
738: \label{Roftau_Rmax}
739: \tau &=& \frac{R_\text{max}}{2 M} \left( \sqrt{\frac{R}{2 M} \left(1 -
740: \frac{R}{R_\text{max}}\right)} \right.+ \nonumber \\
741: &\quad& \quad \qquad \left. \sqrt{\frac{R_\text{max}}{ 2 M }}
742: \arccos{\sqrt{\frac{R}{R_\text{max}}}} \right),
743: \end{eqnarray}
744: where $R = R_\text{max}$ is the position at $\tau = 0$ of a freely
745: falling observer starting at rest, and $M$ is the mass of the black hole.
746: To transform from the $R_\text{max}$ radial coordinate to the
747: isotropic radial coordinate $r$ we use
748: \begin{equation}
749: R_{\text{max}}({r}) = \frac{\left(M + 2 {r}\right)^2}{4 {r}} .
750: \end{equation}
751: Then the ${r}{r}$-component of the metric in isotropic coordinates
752: is given by
753: \begin{equation}
754: \label{eq:novikovg}
755: g_{{r}{r}} = \psi({r})^4
756: \left(\frac{\partial R}{\partial R_\text{max}} \right)^2 ,
757: \end{equation}
758: where
759: \begin{eqnarray}
760: \frac{\partial R}{\partial R_\text{max}} &=& \frac{3}{2} - \frac{1}{2}
761: \frac{R}{R_{\text{max}}} \nonumber + \\
762: &\quad& \frac{3}{2} \sqrt{ \frac{R_{\text{max}}}{R}
763: - 1}\; \arccos{\left(\sqrt{\frac{R}{R_{\text{max}}}}\;\right)}
764: \end{eqnarray}
765: and $R$ is computed from Eq.~(\ref{Roftau_Rmax}).
766: Here $\psi({r}) = 1 + \frac{M}{2 {r}}$.
767: Note that there is a typo in \cite{Bruegmann96},
768: Eq.~(16).
769:
770: Analytically, the metric becomes infinite at time
771: \begin{equation}
772: t_\text{crash} = \pi M,
773: \end{equation}
774: which is the time for an observer that starts from rest at the horizon
775: to reach the Schwarzschild singularity.
776: This results in a crash in the numerical computations. We run this
777: test on a 3D grid in the so-called cartoon mode
778: \cite{Alcubierre99a}, because the spherical symmetry of the problem
779: means that we can do a computation using information only on a
780: line passing trough the center of the black hole. The computational
781: domain is $z \in [0, 80M ], x=0, y=0$. We use second order accurate
782: finite differencing together with an iterative Crank-Nicholson scheme
783: with a Courant factor of $0.25$ for evolution. Since we never run
784: longer than $t_\text{crash} = \pi M$, the outer boundary at $80M$ has
785: no effect on the black hole located at the origin.
786:
787: In order to compute the order of convergence,
788: \begin{equation}
789: \sigma = \log_2 \left| \frac{f_{4 h} - f_{2 h}}{f_{2 h} - f_h} \right| ,
790: \end{equation}
791: we run the test at the resolutions $h=0.01M$, $2h=0.02M$, and
792: $4h=0.04M$. We compute $\sigma$ on each grid-point present in the
793: coarsest run, and $f_{h}$ is the value of the quantity under
794: consideration (here the metric component $g_{zz}$) for resolution $h$.
795: Our grid-points are not in the same places for all three resolutions,
796: because we always stagger the puncture between two grid-points. This
797: means that we must interpolate to get the values at all the coarse
798: grid-points, for which we use 4th order Lagrange interpolation.
799: Fig.~\ref{fig_sigma} shows $\sigma$ for the AA system
800: at time $t=\frac{\pi}{2}M$ and
801: time $t=3.0M$ in the region close to the black hole where the metric
802: deviates the most from the flat conformal metric. We see that we have
803: second order convergence close to the black hole in both cases. The
804: spikes in the curve at later times are due to the fact that the curves
805: of $g_{zz}$ for the 3 resolutions cross.
806:
807:
808:
809:
810:
811: \subsection{1+log lapse, gamma driver shift}
812:
813: We have implemented 1+log lapse and gamma driver shift in our
814: code. This gauge choice makes the BSSN system stable for more than
815: $1000M$ for certain single black hole runs~\cite{Alcubierre02a}. While
816: the geometric motivation of this gauge choice, namely singularity
817: avoidance and reduction of slice-stretching, should be valid for any
818: evolution system, it is unclear whether the AA system will be as
819: stable as BSSN or unstable in this test case. In particular, note that
820: the gamma driver shift is designed to control the evolution of one of
821: the BSSN variables, $\tilde\Gamma^i$.
822: We have not implemented proper outer boundary conditions for the AA
823: system, so in our simulations we have waves coming in from the outer
824: boundaries. We use the same parameters as for the geodesic runs, i.e.\
825: the domain in cartoon mode is $z \in [0, 80M ], x=0, y=0$,
826: the resolution is $ h = 0.01M$, and we use a Courant factor of $0.25$.
827:
828: Our AA run crashes at around $20 M$. Very steep gradients appear near the
829: black hole and they eventually kill the run. The reason for this seems
830: to be that the 1+log lapse depends on the trace of the physical
831: extrinsic curvature. Since we evolve the conformal extrinsic curvature,
832: finite difference errors are enlarged by a factor of $\psi^4$ when
833: computing the physical extrinsic curvature, and these account for the
834: differences in the lapse evolution between BSSN and AA runs. When
835: using the AA system, the lapse drops very fast, leaving a sharp
836: gradient between the frozen region and the region where the fields can
837: evolve. The gradients in the physical variables that kill the run
838: appear where the frozen region meets the dynamic region. In the BSSN
839: system this gradient is more shallow and we do not see this effect.
840:
841: This demonstrates that, as expected, a gauge choice that leads to
842: numerically stable evolutions with one evolution system may well be
843: unstable for others.
844:
845:
846:
847:
848: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
849: \section{Discussion}
850: \label{secDiscuss}
851: \label{Discussion}
852:
853: We have investigated the numerical properties of the previously
854: untested AA evolution system by a variety of numerical experiments,
855: which for comparison we also performed for the ADM and BSSN systems.
856: Since one aspect of the test suite put forward in
857: \cite{Alcubierre2003:mexico-I} is that different numerical
858: implementations of one and the same evolution system can be compared,
859: let us point out that our results for ADM and BSSN agree with
860: \cite{Alcubierre2003:mexico-I} for the specific results shown there (see also
861: \cite{MayaBSSN_web}).
862:
863: However, if the Courant factor in the robust stability test is lowered
864: from $0.5$ (as suggested in~\cite{Alcubierre2003:mexico-I}) to $0.25$
865: we find that the stability of ADM dramatically improves (compare
866: Figs.~\ref{fig_admrobust} and \ref{fig_admrobust025}), while the
867: stability properties of AA remain unchanged. This implies
868: that the Courant factor used in \cite{Alcubierre2003:mexico-I} is too
869: large for the ADM system causing immediate exponential growth, instead of the
870: initially linear growth expected for a weakly hyperbolic system. This
871: confirms a similar observation already described
872: in~\cite{Bona:2003qn}. We use iterative Crank-Nicholson in all
873: our simulations, but~\cite{Bona:2003qn} also point out that
874: dissipation can mask the linear growth expected for
875: ADM. So one should repeat these experiments with a less dissipative
876: scheme like third order Runge-Kutta to look for linear growth in ADM,
877: but also in the AA systems, see Sec.~\ref{sec:hyper}.
878:
879: One property of BSSN that has not been explicitly stated
880: in~\cite{Alcubierre2003:mexico-I} is its drastic failure on periodic
881: domains. This observation is consistent with, and actually
882: strengthens, the notion that BSSN performs well because it is able to
883: propagate modes off the grid~\cite{Alcubierre99e}, in particular for
884: isolated systems with radiative boundary conditions. It will be
885: interesting to see how AA, ADM, and BSSN perform on a gauge wave test
886: with outer boundary conditions that let the gauge wave propagate away
887: from the domain.
888:
889: Our findings for the AA system can be summarized as follows.
890: In the robust stability test, the AA system is stable for a long time,
891: but eventually does go unstable. In this case the AA system does not do
892: as well as the BSSN system. For gauge waves in periodic geometry, the
893: AA system is unstable, but the runs last much longer than the
894: corresponding BSSN runs. The linear wave test
895: shows that the AA system produces a larger drift in the phase than the
896: ADM and BSSN systems, causing the linear wave to propagate faster
897: compared to the analytical solution in the AA evolution than in the
898: other two evolutions.
899:
900: In the black hole runs, AA does as well as ADM and BSSN for
901: geodesic slicing runs, but fails when trying to use gauges that makes
902: BSSN runs long term stable. There are two distinct aspects of this
903: gauge choice. While the gauge is appropriate geometrically
904: (singularity avoiding and countering slice-stretching), there is no reason
905: to expect numerical stability for the AA system. Since it was
906: non-trivial to find a numerically stable gauge choice for BSSN, it is
907: not too surprising that additional work is required to find a gauge
908: choice for the AA system that allows long run times for a single black
909: hole, if indeed such a choice exists.
910:
911:
912: We have found that the tests published in
913: \cite{Alcubierre2003:mexico-I} are helpful in exploring the
914: properties of the AA system, but also that, not unexpectedly, these
915: tests cannot by themselves determine whether an evolution system is
916: worth exploring further for a particular physical system, say black
917: hole evolutions. Note that one should expect that for different
918: physical initial data sets different evolution systems are optimal,
919: see e.g.~\cite{Lindblom:2002et}. Since the relationship between
920: choice of evolution system, gauge choice, outer boundary conditions,
921: and the physical properties of the problem we are trying to simulate
922: is complicated, it is not clear that sufficient understanding of the
923: numerics needed to evolve binary black holes can be gained by singling
924: out for example the evolution system and ignoring the other issues
925: involved. Ultimately, if one wants to evolve black holes, one should
926: evolve black holes.
927:
928: While the AA system has had some success in our numerical
929: experiments, tests with black holes, proper outer boundaries, and
930: appropriate gauge choices are needed to determine if the AA system has
931: a future in numerical relativity.
932:
933:
934:
935:
936: \begin{acknowledgments}
937: We would like to thank A. Alekseenko and D. Arnold for helpful
938: discussions.
939: We acknowledge the support of the Center for Gravitational Wave Physics
940: funded by the National Science Foundation under Cooperative Agreement
941: PHY-01-14375. This work was also supported by NSF grants PHY-02-18750
942: and PHY-98-00973.
943: \end{acknowledgments}
944:
945: \bibliography{references,AA}
946:
947:
948: \newpage
949:
950:
951:
952:
953:
954:
955: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
956: % begin FIGURES
957: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
958:
959: \begin{figure*}[ht]
960: \begin{center}
961: \epsfig{file=ADMrobust.ps, width=15cm}
962: \end{center}
963: \caption[]{\label{fig_admrobust} Robust stability test
964: for the ADM system. The legend is: solid: $\rho=1$, dotted:
965: $\rho=2$, dash-dot: $\rho=4$, $L_\infty(\text{ham})$ is shown both
966: as a function of time (left) and iterations (right). Note the
967: almost perfect alignment when plotting versus the number iterations,
968: which is the signature of a grid mode.}
969: \end{figure*}
970:
971:
972:
973: \begin{figure*}[ht]
974: \begin{center}
975: \epsfig{file=ADMrobust025.ps, width=15cm}
976: \end{center}
977: \caption[]{\label{fig_admrobust025} Robust stability test for the ADM
978: system, with Courant factor $0.25$. The legend is: solid: $\rho=1$, dotted:
979: $\rho=2$, dash-dot: $\rho=4$. We show $L_\infty(\text{ham})$ on the
980: left and $L_\infty(\text{tr$K$})$ on the right. The exponentially
981: growing mode seen in Fig.~\ref{fig_admrobust} is not present.}
982: \end{figure*}
983:
984: \begin{figure*}[ht]
985: \begin{center}
986: \epsfig{file=bssnrobust.ps, width=15cm}
987: \end{center}
988: \caption[]{\label{fig_bssnrobust} Robust stability test
989: for the BSSN system. The legend is: solid: $\rho=1$, dotted:
990: $\rho=2$, dash-dot: $\rho=4$.}
991: \end{figure*}
992:
993:
994: \begin{figure*}[ht]
995: \begin{center}
996: \epsfig{file=AArobust.ps, width=15cm}
997: \end{center}
998: \caption[]{\label{fig_AArobust} Robust stability for the
999: AA system, run until 1000 crossing times.
1000: The legend is: solid: $\rho=1$, dotted:
1001: $\rho=2$, dash-dot: $\rho=4$.}
1002: \end{figure*}
1003:
1004: \begin{figure*}[ht]
1005: \begin{center}
1006: \epsfig{file=AArobustlong.ps, width=15cm}
1007: \end{center}
1008: \caption[]{\label{fig_AArobustlong} Robust stability for
1009: the AA system, run until 10000 crossing times. The legend is:
1010: solid: $\rho=1$, dotted: $\rho=2$, dash-dot: $\rho=4$. We see that at
1011: late times the constraint violations are growing, which will
1012: probably cause a crash if we evolve long enough.}
1013: \end{figure*}
1014:
1015: \begin{figure*}[ht]
1016: \begin{center}
1017: \epsfig{file=AAgauge.ps, width=15cm}
1018: \end{center}
1019: \caption[]{\label{fig_AAgauge} Gauge stability test for
1020: the AA system with a wave amplitude $A=0.1$. The legend is:
1021: solid: $\rho=1$, dotted: $\rho=2$, dash-dot: $\rho=4$. As expected,
1022: the metric converges at early times, but there is exponential growth
1023: and the run crashes after about 100 crossing times.}
1024: \end{figure*}
1025:
1026:
1027: \begin{figure*}[ht]
1028: \begin{center}
1029: \epsfig{file=AAgaugeA001.ps, width=15cm}
1030: \end{center}
1031: \caption[]{\label{fig_AAgaugeA001} Gauge stability test
1032: for the AA system with a wave amplitude $A=0.01$. The legend is:
1033: solid: $\rho=1$, dotted: $\rho=2$, dash-dot: $\rho=4$.
1034: We see that with this smaller amplitude the runs last 10 times longer but
1035: still crash eventually.}
1036: \end{figure*}
1037:
1038: \begin{figure*}[ht]
1039: \begin{center}
1040: \epsfig{file=bssngauge.ps, width=15cm}
1041: \end{center}
1042: \caption[]{\label{fig_bssngauge} Gauge stability test
1043: for the BSSN system with a wave amplitude $A=0.01$. The legend is:
1044: solid: $\rho=1$, dotted: $\rho=2$, dash-dot: $\rho=4$. Both the
1045: metric and the extrinsic curvature tensor become non-convergent in a
1046: short time.}
1047: \end{figure*}
1048:
1049: \begin{figure*}[ht]
1050: \begin{center}
1051: \epsfig{file=linearcomp.ps, width=15cm}
1052: \end{center}
1053: \caption[]{\label{fig_linearcomp} Linear wave stability test for (from
1054: left to right) the AA system, the BSSN system, and the ADM system. We
1055: compare the numerical wave at 100 crossing
1056: times to the analytical solution. The legend is: solid: analytical
1057: solution, dotted: $\rho=1$, dash-dot: $\rho=2$, dash: $\rho=4$.}
1058: \end{figure*}
1059:
1060: \begin{figure*}[ht]
1061: \begin{center}
1062: \epsfig{file=linearnorm.ps, width=15cm}
1063: \end{center}
1064: \caption[]{\label{fig_linearnorm} Linear wave stability test for (from
1065: left to right) the AA system, the BSSN system, and the ADM system. We
1066: show the $L_2$ norm of the difference between the analytical and
1067: numerical values at different crossing times. The legend is: solid:
1068: $\rho=1$, dotted: $\rho=2$, dash-dot: $\rho=4$.}
1069: \end{figure*}
1070:
1071: \begin{figure*}[ht]
1072: \begin{center}
1073: \epsfig{file=sigma.ps,width=15cm}
1074: \end{center}
1075: \caption[]{\label{fig_sigma}
1076: Geodesic slicing of Schwarzschild. Shown is the order of
1077: convergence $\sigma$ of the conformal metric component $\bar g_{zz}$
1078: for the AA system in the computational domain near the black hole
1079: for $t=\frac{\pi}{2}M$ and $t= 3.0M$. We observe second order
1080: convergence near the black hole. }
1081: \end{figure*}
1082:
1083: \begin{figure*}[ht]
1084: \begin{center}
1085: \epsfig{file=crash.ps}
1086: \end{center}
1087: \caption[]{\label{fig_crash}
1088: Geodesic slicing of Schwarzschild. The first panel shows the
1089: analytical conformal metric component $\bar g_{zz}$ as a
1090: function of $z$ in the inner region of the domain (same region as in
1091: Fig.~\ref{fig_sigma}). The analytical solution is shown at 3 different
1092: times: $t=1.55$ (solid), $t=3.0$ (dotted) and $t=3.125$
1093: (dash-dot). The other 3 panels show the absolute value of the relative
1094: difference of the numerical and analytical solutions at the 3
1095: different times. In these plots the AA solution is solid, ADM is
1096: dotted, and BSSN is dash-dot.}
1097: \end{figure*}
1098:
1099:
1100: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1101: \end{document}
1102:
1103:
1104: