1: \documentclass[aps,prd,amssymb,eqsecnum,showpacs] {revtex4}
2: \usepackage{graphicx}
3: \usepackage{psfrag}
4:
5: \begin{document}
6:
7: \title{Strategies for the Characteristic Extraction of Gravitational Waveforms}
8:
9:
10: \author{M.~C. Babiuc${}^{1}$, N.~T. Bishop${}^{2}$, B. Szil\'{a}gyi${}^{3,4}$
11: and J. Winicour${}^{3,5}$
12: }
13: \affiliation{
14: ${}^{1}$ Department of Physics \\
15: Marshall University, Huntington, WV 25755, USA \\
16: ${}^{2}$Department of Mathematical Sciences,
17: University of South Africa, Unisa 0003, South Africa,\\
18: ${}^{3}$ Max-Planck-Institut f\" ur
19: Gravitationsphysik, Albert-Einstein-Institut,
20: 14476 Golm, Germany \\
21: ${}^{4}$ Theoretical Astrophysics, California Institute of Technology \\
22: Pasadena, CA 91125, USA \\
23: ${}^{5}$ Department of Physics and Astronomy \\
24: University of Pittsburgh, Pittsburgh, PA 15260, USA
25: }
26:
27: \begin{abstract}
28:
29: We develop, test and compare new numerical and geometrical methods for improving
30: the accuracy of extracting waveforms using characteristic evolution. The new
31: numerical method involves use of circular boundaries to the stereographic grid
32: patches which cover the spherical cross-sections of the outgoing null cones. We
33: show how an angular version of numerical dissipation can be introduced into the
34: characteristic code to damp the high frequency error arising form the irregular
35: way the circular patch boundary cuts through the grid. The new geometric method
36: involves use of the Weyl tensor component $\Psi_4$ to extract the waveform as
37: opposed to the original approach via the Bondi news function. We develop the
38: necessary analytic and computational formula to compute the $O(1/r)$
39: radiative part of $\Psi_4$ in terms of a conformally compactified treatment of
40: null infinity. These methods are compared and calibrated in test problems based
41: upon linearized waves.
42:
43: \end{abstract}
44:
45: \pacs{04.20Ex, 04.25Dm, 04.25Nx, 04.70Bw}
46:
47: \maketitle
48:
49: \section{Introduction}
50:
51: The unambiguous geometric description of gravitational radiation in curved
52: spacetimes traces back to the work of Bondi~\cite{bondi} et al,
53: Sachs~\cite{sachs} and Penrose~\cite{Penrose}. By formulating asymptotic
54: flatness in terms of characteristic hypersurfaces extending to infinity, they
55: were able to reconstruct, in a nonlinear geometric setting, the basic properties
56: of gravitational waves which had been developed in linearized theory on a
57: Minkowski background. The major new nonlinear features were the Bondi mass and
58: news function, and the mass loss formula relating them. This approach has been
59: implemented as a characteristic evolution code~\cite{isaac,highp} which computes
60: the radiation field at infinity by using a Penrose compactification of the
61: space-time. The code computes the gravitational radiation reaching infinity in
62: terms of boundary data supplied on an inner worldtube. This has timely
63: application to the important astrophysical problem of the inspiral and merger of
64: a binary black hole. Several Cauchy codes, using an artificial outer boundary
65: condition, are now able to simulate this binary problem. By using the data on a
66: worldtube carved out of these binary black hole spacetimes obtained by Cauchy
67: evolution, the characteristic code can supply the resulting waveform at
68: infinity. In this work, we develop and test new methods designed to enhance the
69: accuracy of this approach to computing gravitational waveforms, which has been
70: called {\it Cauchy-characteristic extraction} (CCE)~\cite{cce}.
71:
72: The Cauchy codes presently being applied to the binary black hole problem
73: introduce an artificial outer boundary, where some boundary condition must be
74: employed. The choice of the proper boundary condition for an isolated radiating
75: system is a global problem, which can only be treated exactly by an extension of
76: the solution to infinity, e.g. by conformal compactification. The most elegant such
77: approach is the extension of the Cauchy problem to future null infinity ${\cal
78: I}^+$ by means of a hyperboloidal time foliation~\cite{hypb}
79: (see~\cite{Hhprbloid,joerg} for reviews of progress in this direction). Another
80: approach is to extend the solution to ${\cal I}^+$ by matching the interior
81: Cauchy evolution to an exterior characteristic evolution, i.e.
82: Cauchy-characteristic matching (CCM)~\cite{ccm}. (see~\cite{livccm} for a review). CCM has
83: been applied successfully to gravitational wave computations in the linear
84: regime~\cite{harm} but has not yet been extended to the nonlinear binary black
85: hole problem.
86:
87: When an artificial finite outer boundary is introduced there are two broad
88: sources of error:
89: \begin{itemize}
90:
91: \item The outer boundary condition
92:
93: \item Waveform extraction at an inner worldtube.
94: \end{itemize}
95:
96: The first source of error stems from the outer boundary condition, which must
97: lead to a well-posed constraint-preserving initial-boundary value problem. This
98: has not yet been fully established for any of the present black hole codes. But,
99: even were such boundary conditions implemented, the correct boundary data must
100: be prescribed. However, this boundary data can only be exactly determined, in
101: general, by extending the solution to infinity~\cite{gk}. Otherwise, the best
102: that can be done is to impose a boundary condition for which {\it homogeneous}
103: boundary data, i.e. zero boundary values, is a good approximation. One proposal
104: of this type~\cite{friednag} is a boundary condition that requires the
105: Newman-Penrose~\cite{NP} Weyl tensor component $\Psi_0$ to vanish. In the limit that
106: the outer boundary goes to infinity this outer boundary condition becomes exact.
107: In the present state of the art of black hole simulations, this approach comes
108: closest to a satisfactory treatment of the outer boundary~\cite{psiocit}.
109:
110: The second source of error arises from waveform extraction at an inner
111: worldtube, which must be well inside the outer boundary in order to isolate it
112: from errors introduced by the boundary condition. There the waveform is
113: typically extracted by a perturbative scheme based upon the introduction of a
114: background Schwarzschild spacetime. This has been carried out using the
115: Regge-Wheeler-Zerilli ~\cite{regwh,zeril} treatment of the perturbed metric, as
116: reviewed in~\cite{nagrez}, and also by calculating the Newman-Penrose Weyl
117: component $\Psi_4$, as first done for the binary black hole problem
118: in~\cite{bakcamluo,pretlet,camluomarzlo,bakcenchokopvme}. In this approach,
119: errors arise from the finite size of the extraction worldtube, from
120: nonlinearities and from gauge ambiguities involved in the arbitrary introduction
121: of a background metric. The gauge ambiguities might seem less severe in the case
122: of $\Psi_4$ (vs metric) extraction, but there are still delicate problems
123: associated with the choices of a preferred null tetrad and preferred worldlines
124: along which to measure the waveform (see ~\cite{lehnmor} for an analysis).
125:
126: Cauchy-characteristic extraction, which is one of the pieces of the CCM
127: strategy, offers a means to avoid this error introduced by extraction at a
128: finite worldtube. In CCE, the inner worldtube data supplied by the Cauchy
129: evolution is used as boundary data for a characteristic evolution to future null
130: infinity, where the waveform can be unambiguously computed by geometric methods.
131: By itself, CCE does not use the characteristic evolution to inject outer
132: boundary data for the Cauchy evolution, which can be a source of instability in
133: full CCM. Highly nonlinear tests in black hole spacetimes~\cite{highp} have
134: shown that characteristic evolution is a stable procedure which provides the
135: geometry in the neighborhood of null infinity up to numerical error; and tests
136: in the perturbative regime~\cite{babiuc05} show that CCE compares favorably with
137: Zerilli extraction and has advantages at small extraction radii. However, in
138: astrophysically realistic cases which require high resolution, such as the
139: inspiral of matter into a black hole~\cite{partbh}, this error has been a
140: troublesome factor in the postprocessing of the numerical solution which is
141: necessary to compute the asymptotic quantities determining the Bondi news
142: function.
143:
144: There are two distinct ways, geometric and numerical, that the accuracy of this
145: calculation of the gravitational waveform might be improved. In the geometrical
146: category, one option is to compute $\Psi_4$ instead of the news function as the
147: primary description of the waveform. We discuss this in
148: Sec.~\ref{sec:waveforms}, where we develop the extensive formulae necessary to
149: compute the asymptotic $O(1/r)$ part of $\Psi_4$, i.e.
150: $\Psi_4^0=\lim_{ r\rightarrow \infty }r\Psi_4$, which governs the radiation.
151:
152: In the numerical category, some standard methods for improving accuracy, such
153: as higher order finite difference approximations, would be easy to implement
154: whereas others, such as adaptive mesh refinement, have only been tackled for 1D
155: characteristic codes~\cite{pretlehn}. But beyond these methods, a major source
156: of error in characteristic evolution arises from the intergrid interpolations
157: arising from the multiple patches necessary to coordinatize smoothly the
158: spherical cross-sections of the outgoing null hypersurfaces. The development of
159: grids smoothly covering the sphere has had a long history in computational
160: meteorology that has led to two distinct approaches: (i) the stereographic
161: approach in which the sphere is covered by two overlapping patches obtained by
162: stereographic projection about the North and South poles~\cite{browning}; and
163: (ii) the cubed-sphere approach in which the sphere is covered by the 6 patches
164: obtained by a projection of the faces of a circumscribed cube~\cite{ronchi}.
165: Recently, the cubed-sphere approach has received much attention because the
166: simple structure of its shared boundaries allows a highly scalable algorithm for
167: parallel architectures. A discussion of the advantages of each of these methods
168: and a comparison of their performance in a standard fluid testbed are given
169: in~\cite{browning}. In numerical relativity, the stereographic method has been
170: reinvented in the context of the characteristic evolution problem~\cite{eth};
171: and the cubed-sphere method reinvented in the context of building an apparent
172: horizon finder~\cite{Thornburgah}. The characteristic evolution code was first
173: developed using two square stereographic patches, each overlapping the equator.
174: We consider here a modification, based upon the approach advocated
175: in~\cite{browning}, which retains the original stereographic patch structure but
176: shrinks the overlap region by masking a circular boundary near the equator.
177: Recently, the cubed-sphere method has also been developed for application to
178: characteristic evolution~\cite{reisswig,roberto}.
179:
180: These geometric and numerical considerations lead to four options for improving
181: CCE:
182:
183: \begin{itemize}
184:
185: \item Computation of the news function using circular stereographic patches
186:
187: \item Computation of the Weyl tensor using circular stereographic patches
188:
189: \item Computation of the news function using the cubed-sphere patching
190:
191: \item Computation of the Weyl tensor using cubed-sphere patching
192:
193: \end{itemize}
194:
195: We compare these options here in the context of model problems designed to test
196: their application to CCE. Because the cubed-sphere approach requires further
197: code development to be applied to CCE, in Sec.~\ref{sec:stests} we present a
198: test based upon the propagation of a wave on the sphere to provide a preliminary
199: comparison with the stereographic approach. The test compares their accuracy in
200: calculating the angular derivatives required in the news and Weyl tensor
201: extraction algorithms. In Sec.~\ref{sec:extraction}, we next present tests of
202: CCE which compare the news and Weyl tensor approaches in a linearized
203: gravitational wave test problem.
204:
205: The development of finite-difference evolution algorithms, which was largely
206: motivated by application to computational fluid dynamics (CFD). It has utilized
207: the method of lines, where a 3-dimensional spatial domain is discretized to
208: yield a set of coupled ordinary differential equations in time for the grid
209: values, which are then integrated, e.g. by a Runge-Kutta procedure. This $3+1$
210: approach is not applicable to the $2+1+1$ format of characteristic evolution
211: considered here, in which the discretization of a 2-dimensional spherical set of
212: characteristics leads to coupled 2-dimensional partial differential equations in
213: the plane spanned by the outgoing and ingoing characteristics. This $2+1+1$
214: approach is natural to general relativity since the characteristics (light rays)
215: are fundamental to describing the dynamical geometry of space-time. It would be
216: impractical in CFD in which the characteristics have a complicated dynamic
217: relation (determined by equations of state) to the fixed Euclidean geometry. As
218: a result, characteristic evolution algorithms were developed only recently in the
219: context of general relativity and there has been relatively little analysis of
220: their computational properties. In particular, for CFD or any symmetric
221: hyperbolic system, numerical dissipation can be added in the standard
222: Kreiss-Oliger form~\cite{kreisoligd}. One of the main results of this paper is
223: to show how analogous dissipation can be successfully applied in a $2+1+1$
224: format. In the original version of the PITT code, which used square
225: stereographic patches with boundaries aligned with the grid, numerical
226: dissipation was only introduced in the radial direction~\cite{luisdis}. This was
227: sufficient to establish numerical stability. In the new version of the code with
228: circular stereographic patches, whose boundaries do not fit regularly on the
229: stereographic grid, numerical dissipation is necessary to control the high
230: frequency error introduced by the intergrid interpolations, as previously noted
231: in the treatment of a fluid problem using circular stereographic
232: patches~\cite{browning}. We begin with a brief review of the formalism
233: underlying the characteristic evolution code in Sec.~\ref{sec:chform} and show
234: how the essential new feature of angular dissipation can be incorporated.
235:
236: The two spherical grid methods, stereographic and cubed sphere, are briefly
237: described in Sec.~\ref{sec:patches}. We present the test results in
238: Secs.~\ref{sec:stests} and \ref{sec:extraction} and we summarize our conclusions
239: in Sec.~\ref{sec:concl}.
240:
241: \section{Characteristic Formalism}
242: \label{sec:chform}
243:
244: The characteristic formalism is based upon a family of outgoing null
245: hypersurfaces, emanating from some inner worldtube, which extend to
246: infinity where they foliate ${\cal I}^+$ into spherical slices.
247: We let $u$ label these hypersurfaces, $x^A$ $(A=2,3)$ be angular coordinates
248: which label the null rays and $r$ be a surface area coordinate. In the
249: resulting $x^\alpha=(u,r,x^A)$ coordinates, the metric takes the Bondi-Sachs
250: form~\cite{bondi,sachs}
251: \begin{eqnarray}
252: ds^2 & = & -\left(e^{2\beta}\frac{V}{r} -r^2h_{AB}U^AU^B\right)du^2
253: -2e^{2\beta}dudr -2r^2 h_{AB}U^Bdudx^A \nonumber \\
254: & + & r^2h_{AB}dx^Adx^B,
255: \label{eq:bmet}
256: \end{eqnarray}
257: where $h^{AB}h_{BC}=\delta^A_C$ and
258: $det(h_{AB})=det(q_{AB})$, with $q_{AB}$ a unit sphere metric. In
259: analyzing the Einstein equations, we also use the intermediate variable
260: \begin{equation}
261: Q_A = r^2 e^{-2\,\beta} h_{AB} U^B_{,r}.
262: \end{equation}
263:
264: The code introduces an auxiliary unit sphere metric $q_{AB}$, with
265: associated complex dyad $q_A$ satisfying
266: $ q_{AB} =\frac{1}{2}\left(q_A \bar q_B+\bar q_Aq_B\right)$.
267: For a general Bondi-Sachs metric,
268: $h_{AB}$ can then be represented by its dyad component $J=h_{AB}q^Aq^B
269: /2$, with the spherically symmetric case characterized by $J=0$. The
270: full nonlinear $h_{AB}$ is uniquely determined by $J$, since the
271: determinant condition implies that the remaining dyad component
272: $K=h_{AB}q^A \bar q^B /2$ satisfies $1=K^2-J\bar J$. We also introduce
273: spin-weighted fields $U=U^Aq_A$ and $Q=Q_Aq^A$, as well as the (complex
274: differential) operators $\eth$ and $\bar \eth$~\cite{newp}.
275: Refer to {}~\cite{eth,cce} for further details.
276:
277: In this formalism, the Einstein equations $G_{\mu\nu}=0$ decompose into
278: hypersurface equations, evolution equations and conservation conditions on the
279: inner worldtube. As described in more detail in~\cite{newt,nullinf}, the
280: hypersurface equations take the form
281: \begin{eqnarray}
282: \beta_{,r} &=& N_\beta,
283: \label{eq:beta} \\
284: U_{,r} &=& r^{-2}e^{2\beta}Q +N_U,
285: \label{eq:wua} \\
286: (r^2 Q)_{,r} &=& -r^2 (\bar \eth J + \eth K)_{,r}
287: +2r^4\eth \left(r^{-2}\beta\right)_{,r} + N_Q,
288: \label{eq:wq} \\
289: V_{,r} &=& \frac{1}{2} e^{2\beta}{\cal R}
290: - e^{\beta} \eth \bar \eth e^{\beta}
291: + \frac{1}{4} r^{-2} \left(r^4
292: \left(\eth \bar U +\bar \eth U \right)
293: \right)_{,r} + N_W,
294: \label{eq:ww}
295: \end{eqnarray}
296: where~\cite{eth}
297: \begin{equation}
298: {\cal R} =2 K - \eth \bar \eth K + \frac{1}{2}(\bar \eth^2 J + \eth^2 \bar J)
299: +\frac{1}{4K}(\bar \eth \bar J \eth J - \bar \eth J \eth \bar J)
300: \label{eq:calR}
301: \end{equation}
302: is the curvature scalar of the 2-metric $h_{AB}$.
303: The evolution equation takes the form
304: \begin{eqnarray}
305: && 2 \left(rJ\right)_{,ur}
306: - \left(r^{-1}V\left(rJ\right)_{,r}\right)_{,r} = \nonumber \\
307: && -r^{-1} \left(r^2\eth U\right)_{,r}
308: + 2 r^{-1} e^{\beta} \eth^2 e^{\beta}- \left(r^{-1} V \right)_{,r} J
309: + N_J,
310: \label{eq:wev}
311: \end{eqnarray}
312: where, $N_\beta$, $N_U$, $N_Q$, $N_W$ and $N_J$ are nonlinear terms which
313: vanish for spherical symmetry. Expressions for these terms as complex
314: spin-weighted fields and a discussion of the conservation conditions are given
315: in~\cite{cce}.
316:
317: The characteristic evolution code implements this formalism as an explicit
318: finite difference scheme. In this paper we use second order accurate finite
319: differences and we reduce all angular derivatives to first order by the
320: introduction of auxiliary variables, as described in~\cite{gomezfo}
321:
322: \subsection{Angular dissipation}
323: \label{sec:angdiss}
324:
325: It is a feature of the composite mesh technique that numerical dissipation is
326: necessary to stabilize the error introduced by intergrid interpolations. In the
327: case of a square stereographic patch, whose boundary aligns with the grid lines,
328: the dissipation built into the characteristic radial integration scheme is
329: sufficient for this purpose~\cite{luisdis}. However, because a circular
330: boundary fits into a stereographic grid in an irregular way, angular
331: dissipation is also necessary in order to suppress the resulting high frequency
332: error introduced by the interpolations between stereographic patches.
333:
334: We accomplish this by modifying the evolution equation (\ref{eq:wev}) as follows. In the
335: code, (\ref{eq:wev}) is expressed in terms of a compactified radial coordinate
336: $x=r/(R+r)$, where $R$ is an adjustable scale parameter and ${\cal I}^+$ has
337: finite coordinate value $x=1$. The evolution in retarded time $u$ is carried out in terms of the variable
338: $\Phi=xJ$, which is regular at ${\cal I}^+$. Then the evolution equation
339: (\ref{eq:wev}) takes the form
340: \begin{equation}
341: \partial_u \bigg((1-x) \Phi_{,x} +\Phi \bigg)= S,
342: \label{eq:phiev}
343: \end{equation}
344: where $S$ represents the right hand side terms.
345: We add angular dissipation to the $u$-evolution through the modification
346: \begin{equation}
347: \partial_u \bigg ((1-x) \Phi_{,x} +\Phi \bigg )
348: +\epsilon_{u} h^3 \eth^2 {\cal W}
349: \bar\eth^2 \bigg ((1-x) \Phi_{,x} +\Phi \bigg )
350: = S,
351: \end{equation}
352: where $h$ is the discretization size, $\epsilon_{u}\ge 0$ is an adjustable
353: parameter independent of $h$ and $ {\cal W}$ is a positive weighting function
354: with $ {\cal W}=1$ inside the equator and $ {\cal W}=0$ at the patch boundary.
355: This leads to
356: \begin{equation}
357: \partial_u \bigg( |(1-x) \Phi_{,x} +\Phi|^2\bigg)
358: + 2\epsilon_{u} h^3 \Re\{ \bigg ((1-x) \bar \Phi_{,x} +\bar \Phi \bigg )
359: \eth^2 {\cal W}\bar\eth^2 \bigg ((1-x) \Phi_{,x} +\Phi \bigg ) \}
360: = 2\Re \{\bigg ((1-x) \bar \Phi_{,x} +\bar \Phi \bigg ) S \}.
361: \end{equation}
362: Integration over the unit sphere with solid angle element $d\Omega$ then gives
363: \begin{equation}
364: \partial_u \oint | (1-x) \Phi_{,x} +\Phi|^2 d\Omega
365: +2\epsilon_{u} h^3 \oint
366: {\cal W}|\bar\eth^2 \bigg ((1-x)\Phi_{,x}+\Phi \bigg )|^2 d\Omega
367: =2\Re \oint \bigg ((1-x) \bar \Phi_{,x} +\bar \Phi \bigg ) S d\Omega.
368: \end{equation}
369: Thus the $\epsilon_{u}$-term has the effect of damping high frequency noise as
370: measured by the $L_2$ norm of $(1-x) \Phi_{,x} +\Phi$ over the sphere.
371:
372: Similarly, dissipation is introduced in the radial integration of
373: (\ref{eq:phiev}) through the substitution
374: \begin{equation}
375: \partial_u \bigg ((1-x) \Phi_{,x} +\Phi \bigg ) \rightarrow
376: \partial_u \bigg ((1-x) \Phi_{,x} +\Phi \bigg )
377: +\epsilon_{x} h^3 \eth^2 {\cal W} \bar\eth^2 \Phi_{,u}
378: \end{equation}
379: with $\epsilon_{x}\ge 0$ .
380: Angular dissipation is also introduced in the hypersurface
381: equations through the substitutions
382: \begin{eqnarray}
383: (r^2 Q)_{,r} &\rightarrow& (r^2 Q)_{,r}
384: +\epsilon_Q h^3 \eth \bar \eth {\cal W} \eth\bar \eth r Q \\
385: V_{,r} &\rightarrow & V_{,r}
386: +\epsilon_V h^3 \eth\bar\eth {\cal W} \eth\bar\eth V.
387: \end{eqnarray}
388:
389:
390: \section{Waveforms at ${\cal I}^+$}
391: \label{sec:waveforms}
392:
393: For an analytic treatment of the Penrose compactification of an asymptotically flat
394: space-time, it is simplest to introduce an inverse radial coordinate $\ell=1/r$,
395: so that future null infinity ${\cal I}^+$ is given by $\ell=0$~\cite{tam}. In
396: the resulting $x^\mu=(u,\ell,x^A)$ conformal Bondi coordinates, the physical
397: space-time metric $g_{\mu\nu}$ has the conformal compactification $\hat
398: g_{\mu\nu}=\ell^{2} g_{\mu\nu}$, where $\hat g_{\mu\nu}$ is smooth at ${\cal
399: I}^+$ and, referring to (\ref{eq:bmet}), takes the form
400: \begin{equation}
401: \hat g_{\mu\nu}dx^\mu dx^\nu=
402: -\left(e^{2\beta}V \ell^3 -h_{AB}U^AU^B\right)du^2
403: +2e^{2\beta}dud\ell -2 h_{AB}U^Bdudx^A + h_{AB}dx^Adx^B.
404: \label{eq:lmet}
405: \end{equation}
406: The inverse conformal metric has the non vanishing components
407: $\hat g^{u\ell} =e^{-2\beta}$,
408: $\hat g^{\ell \ell} =e^{-2\beta}\ell^3 V$,
409: $\hat g^{\ell A} =e^{-2\beta}U^A$ and $\hat g^{AB}=h^{AB}$.
410:
411: The Bondi mass, news function and $\Psi_4^0$ (functions of $u$ and $x^A$),
412: which describe the total energy and radiation power, are constructed from the
413: leading coefficients in an expansion of the metric in powers of $\ell$. The
414: requirement of an asymptotically flat vacuum exterior imposes relations between
415: these expansion coefficients. In the $\hat g_{\mu\nu}$ conformal frame, the
416: vacuum gravitational equations are
417: \begin{equation}
418: -\ell^2 \hat G_{\mu\nu} =2\ell (\hat\nabla_\mu \hat\nabla_\nu \ell
419: -\hat g_{\mu\nu} \hat \nabla^\alpha \hat\nabla_\alpha \ell )
420: +3\hat g_{\mu\nu} (\hat\nabla^\alpha \ell) \hat\nabla_\alpha \ell
421: \label{eq:einstein}
422: \end{equation}
423: in terms of the Einstein tensor $\hat G_{\mu\nu}$ and covariant derivative $\hat
424: \nabla_\mu$ associated with $\hat g_{\mu\nu}$. Asymptotic flatness immediately
425: implies that $\hat g^{\ell \ell} =(\hat\nabla^\alpha \ell) \hat\nabla_\alpha
426: \ell =O(\ell)$ so that ${\cal I}^+$ is null. From the trace of
427: (\ref{eq:einstein}), we have
428: \begin{equation}
429: (\hat\nabla^\alpha \ell) \hat\nabla_\alpha \ell=
430: \frac{1}{2}\ell \hat\Theta+O(\ell^2),
431: \end{equation}
432: where
433: \begin{equation}
434: \hat\Theta:=\hat\nabla^\mu \hat\nabla_\mu \ell
435: =e^{-2\beta} \bigg( \partial_\ell (\ell^3 V)+\partial_A U^A \bigg)
436: \label{eq:theta}
437: \end{equation}
438: is smooth at ${\cal I}^+$. In addition, (\ref{eq:einstein}) implies the
439: existence of a smooth trace-free field $\hat\Sigma_{\mu\nu}$ defined by
440: \begin{equation}
441: \ell \hat \Sigma_{\mu\nu} :=
442: \hat \nabla_\mu \hat\nabla_\nu \ell
443: -\frac{1}{4}\hat g_{\mu\nu}\hat\Theta.
444: \label{eq:Sigma}
445: \end{equation}
446: For future reference we introduce an orthonormal null tetrad $(\hat n^\mu, \hat
447: \ell^\mu, \hat m^\mu)$ be such that $\hat n^\mu=\hat \nabla^\mu \ell$ and
448: $\hat \ell^\mu \partial\mu=\partial_\ell$ at ${\cal I}^+$. Note that
449: (\ref{eq:einstein}), (\ref{eq:theta}) and (\ref{eq:Sigma}) imply
450: \begin{equation}
451: \hat m^\nu \hat m^\rho (\hat\Sigma_{\nu\rho}+\frac{1}{2}\hat G_{\nu\rho})=0.
452: \label{eq:einsteinsig}
453: \end{equation}
454: The gravitational waveform depends on the value of $\hat\Sigma_{\mu\nu}$ on
455: ${\cal I}^+$, which in turn depends on the leading terms up to $O(\ell)$ in the
456: expansion of $\hat g_{\mu\nu}$. We thus expand
457: \begin{equation}
458: h_{AB}(u,\ell,x^C)= H_{AB}(u,x^C)+\ell c_{AB}(u,x^C)+O(\ell^2).
459: \end{equation}
460: Further conditions on the asymptotic expansion of the metric can be extracted
461: from (\ref{eq:einstein}). We have
462: \begin{equation}
463: \beta(u,\ell,x^C)=H(u,x^C)+ O(\ell^2)
464: \end{equation}
465: (where the $O(\ell)$ term vanishes),
466: \begin{equation}
467: U^A= L^A+2\ell e^{2H} H^{AB}D_B H+O(\ell^2),
468: \end{equation}
469: and
470: \begin{equation}
471: \ell^2 V= D_A L^A
472: +\ell (e^{2H}{\cal R}/2 +D_A D^A e^{2H})+O(\ell^2),
473: \end{equation}
474: where ${\cal R}$ and $D_A$ are the 2-dimensional curvature scalar and covariant
475: derivative associated with $h_{AB}$. These results combine with (\ref{eq:theta})
476: to give
477: \begin{equation}
478: \hat\Theta =2e^{-2H}D_A L^A +\ell \bigg ({\cal R}
479: + 3 e^{-2H}D_A D^A e^{2H} \bigg )+O(\ell^2).
480: \end{equation}
481: In addition, the requirement that
482: $$
483: \ell ( \hat \Sigma_{AB} -\frac{1}{2}H_{AB} H^{CD}\Sigma_{CD})
484: $$
485: vanishes at ${\cal I}^+$ implies via (\ref{eq:Sigma}) that
486: \begin{equation}
487: 2H_{C (A} D_{B)} L^C+\partial_u H_{AB}
488: - H_{AB} D_{C} L^C =O(\ell).
489: \label{eq:asymHu}
490: \end{equation}
491:
492: The expansion coefficients $H$, $H_{AB}$, $c_{AB}$ and $L^A$ (all functions of
493: $(u,x^A)$) completely determine the radiation field. One can further specialize
494: the Bondi coordinates to be {\em inertial} at ${\cal I}^+$, i.e. have Minkowski
495: form, in which case $H=L^A=0$, $H_{AB}=q_{AB}$ (the unit sphere metric) so that
496: (\ref{eq:asymHu}) is trivially satisfied and the radiation field is determined
497: by $c_{AB}$. However, the characteristic extraction of the waveform is carried
498: out in null coordinates determined by data on the inner worldtube so that this
499: {\em inertial} simplification cannot be assumed.
500:
501: \subsection{Calculation of the news}
502:
503: The following calculation of the Bondi news streamlines the presentation
504: in~\cite{highp} and corrects errors. In order to carry out the
505: calculation in the $\hat g_{\mu\nu}$ computational frame, it is useful to refer
506: to an inertial conformal Bondi frame~\cite{tam} with metric $\tilde
507: g_{\mu\nu}=\Omega^2 g_{\mu\nu} =\omega^2 \hat g_{\mu\nu}$, where
508: $\Omega=\omega\ell$, which satisfies the gauge requirements that
509: $Q_{AB}:={\tilde g}_{AB}|_{{\cal I}^+}=\omega^2 H_{AB}$ is intrinsically a unit
510: sphere metric at ${\cal I^+}$ and that $(\tilde \nabla^{\alpha} \Omega)\tilde
511: \nabla_{\alpha} \Omega=O(\Omega^2)$. (See~\cite{quad} for a discussion of how
512: the news in an arbitrary conformal frame is related to its expression in this
513: inertial Bondi frame.)
514:
515: ${\cal I^+}$ is a null hypersurface with the null vector ${\tilde
516: n}^{\alpha}={\tilde g}^{\alpha \beta}\nabla_{\beta} \Omega |_{\cal I^+}$, or
517: equivalently, ${\hat n}^{\alpha} = {\hat g}^{\alpha\beta}\nabla_{\beta} \ell
518: |_{\cal I^+}=\omega{\tilde n}^{\alpha}$, tangent to its generators. In order to
519: complete a basis for tangent vectors to ${\cal I^+}$, let ${\cal Q}^{\alpha}$ be
520: a complex field tangent to ${\cal I^+}$ satisfying ${\cal Q}^{\alpha} \tilde
521: n_\alpha=0$, ${\tilde g}_{\alpha\beta}{\cal Q}^{\alpha}{\cal Q}^{\beta} |_{\cal
522: I^+}=0$ and ${\tilde g}_{\alpha\beta}{\cal Q}^{\alpha}{\bar {\cal Q}}^{\beta}
523: |_{\cal I^+}=2$. In an inertial conformal Bondi frame, the news function can
524: then be expressed as~\cite{highp}
525: \begin{equation}
526: N=\lim_{\Omega \rightarrow 0}{1\over 2\Omega}{\cal Q}^{\alpha} {\cal Q}^{\beta}
527: {\tilde \nabla}_{\alpha} {\tilde \nabla}_{\beta} \Omega
528: \label{eq:snews}
529: \end{equation}
530: evaluated in the limit of ${\cal I^+}$. (Our conventions are chosen so
531: that the news reduces to Bondi's original expression in
532: the axisymmetric case~\cite{bondi}). In terms of
533: the ${\hat g}_{\alpha\beta}$ frame, with conformal factor $ \ell=\Omega/\omega$,
534: we then have
535: \begin{eqnarray}
536: N=\lim_{\ell \rightarrow 0}{1\over 2}{\cal Q}^{\alpha} {\cal Q}^{\beta} \bigg(
537: {{\hat \nabla}_{\alpha} {\hat \nabla}_{\beta}
538: {\ell} \over {\ell}}
539: -\omega{\hat \nabla}_{\alpha} {\hat \nabla}_{\beta} {1\over \omega}
540: +\frac{1}{\ell \omega}\hat g_{\alpha\beta}
541: (\hat\nabla^\mu \ell) \hat\nabla_\mu \omega \bigg ) \\
542: ={1\over 2}{\cal Q}^{\alpha} {\cal Q}^{\beta} \bigg (
543: \hat \Sigma_{\alpha\beta}
544: -\omega{\hat \nabla}_{\alpha} {\hat \nabla}_{\beta} {1\over \omega}
545: +\frac{1}{\omega}(\partial_\ell \hat g_{\alpha\beta} )
546: (\hat\nabla^\mu \ell) \hat\nabla_\mu \omega
547: \bigg ).
548: \label{eq:lnews}
549: \end{eqnarray}
550: (This corrects an error in equation (30) of~\cite{highp}.) We determine $\omega$
551: on ${\cal I}^+$ in the ${\hat g}_{\alpha\beta}$ frame by solving the elliptic
552: equation governing the conformal transformation of the curvature scalar
553: (\ref{eq:calR}) of the geometry intrinsic to a $u=constant$ cross-section
554: to a unit sphere geometry,
555: \begin{equation}
556: {\cal R}=2(\omega^2+H^{AB}D_A D_B \log \omega).
557: \label{eq:conf}
558: \end{equation}
559: The condition that $(\tilde \nabla^{\alpha} \Omega) \tilde \nabla_{\alpha}
560: \Omega=O(\Omega^2)$ determines the time dependence of $\omega$,
561: \begin{equation}
562: 2{\hat n}^{\alpha} \partial_{\alpha} \log \omega
563: =-e^{-2H}D_AL^A,
564: \label{eq:omegadot}
565: \end{equation}
566: which is used to evolve $\omega$ given a solution of
567: (\ref{eq:conf}) as initial condition.
568:
569: In order to obtain an explicit expression for the news (\ref{eq:lnews}) in the
570: ${\hat g}_{\alpha\beta}$ frame we need to fix the choice of ${\cal Q}^{\beta}$.
571: The freedom ${\cal Q}^{\beta} \rightarrow {\cal Q}^{\beta} + \lambda {\tilde
572: n}^{\beta}$ leaves (\ref{eq:lnews}) invariant but it is important for physical
573: interpretation to choose the spin rotation freedom ${\cal Q}^{\beta} \rightarrow
574: e^{-i\alpha} {\cal Q}^{\beta}$ to satisfy ${\tilde n}^{\alpha}{\tilde
575: \nabla}_{\alpha} {\cal Q}^{\beta}=O(\Omega)$, so that the polarization frame is
576: parallel propagated along the generators of ${\cal I}^+$. This fixes the
577: polarization modes determined by the real and imaginary parts of the news to
578: correspond to those of inertial observers at ${\cal I}^+$.
579:
580: We accomplish this by introducing the dyad
581: decomposition $H^{AB}=(F^A{\bar F}^B+{\bar F}^A F^B)/2$ where
582: \begin{equation}
583: F^A = q^A \sqrt{ \frac{(K+1)}{2 } }
584: -\bar q^A J \sqrt{ 1 \over 2(K+1)} .
585: \end{equation}
586: We set ${\cal Q}^{\beta}=e^{-i\delta}\omega^{-1}F^\beta+\lambda {\tilde
587: n}^{\beta}$, where $F^\alpha:=(0,0,F^A)$. The requirement of an inertial
588: polarization frame, ${\tilde n}^{\alpha}{\tilde \nabla}_{\alpha} {\cal
589: Q}^{\beta}=O(\Omega)$, then determines the time dependence of the phase
590: $\delta$. We obtain, after using (\ref{eq:omegadot}) to eliminate the time
591: derivative of $\omega$,
592: \begin{equation}
593: 2i(\partial_u +L^A\partial_A)\delta = D_A L^A
594: +H_{AC} \bar F^C ( (\partial_u +L^B \partial_B) F^A
595: - F^B \partial_B L^A) .
596: \label{eq:evphase}
597: \end{equation}
598:
599: We can now express the inertial news (\ref{eq:lnews}) in the ${\hat
600: g}_{\alpha\beta}$ frame as
601: \begin{equation}
602: N={1\over 2}e^{-2i\delta}\omega^{-2}F^{\alpha} F^{\beta} \bigg(
603: \hat \Sigma_{\alpha\beta}
604: -\omega{\hat \nabla}_{\alpha} {\hat \nabla}_{\beta} {1\over \omega}
605: +\frac{1}{\omega}(\partial_\ell \hat g_{\alpha\beta} )
606: (\hat\nabla^\mu \ell) \hat\nabla_\mu \omega \bigg).
607: \label{eq:hnews}
608: \end{equation}
609: with $F^{\alpha}=(0,0,F^A)$. An explicit calculation leads to
610: \begin{equation}
611: N={1\over 4}e^{-2i \delta}\omega^{-2}e^{-2H}F^A F^B
612: \{(\partial_u+{\pounds_L})c_{AB}-{1\over 2}c_{AB} D_C L^C
613: +2\omega D_A[\omega^{-2}D_B(\omega e^{2H})]\},
614: \label{eq:news}
615: \end{equation}
616: where $\pounds_L$ denotes the Lie derivative with respect to $L^A$. This
617: corrects a minus sign error in (38) of~\cite{highp}, where spin-weighted
618: expressions for the terms in (\ref{eq:news}) are given.
619:
620: In inertial Bondi coordinates, the expression for the news function reduces to
621: the simple form
622: \begin{equation}
623: N={1\over 4}Q^A Q^B \partial_u c_{AB}.
624: \label{eq:inews}
625: \end{equation}
626: However, the general form (\ref{eq:news}) must be used in the computational
627: coordinates, which is challenging for maintaining accuracy because of the
628: appearance of second angular derivatives of $\omega$.
629:
630: \subsection{Calculation of the Weyl tensor} \label{sec:weyl}
631:
632: Asymptotic flatness implies that the Weyl tensor vanishes at ${\cal I}^+$, i.e.
633: $\hat C_{\mu\nu\rho\sigma}=O(\ell)$ in the $\hat g_{\mu\nu}$ conformal Bondi
634: frame (\ref{eq:lmet}). This is the conformal space version of the peeling
635: property of asymptotically flat spacetimes~\cite{Penrose}. In terms of the
636: orthonormal null tetrad $(\hat n^\mu, \hat \ell^\mu, \hat m^\mu)$, with
637: $\hat n^\mu=\hat \nabla^\mu \ell$ and $\hat \ell^\mu \partial\mu=\partial_\ell$ at ${\cal I}^+$, the radiation is described by the limit
638: \begin{equation}
639: \hat \Psi:=-\frac{1}{2} \lim_{\ell \rightarrow 0}\frac{1}{\ell}
640: \hat n^\mu \hat m^\nu \hat n^\rho \hat m^\sigma \hat C_{\mu\nu\rho\sigma},
641: \label{eq:psi}
642: \end{equation}
643: which corresponds in Newman-Penrose notation to $-(1/2)\bar \psi_4^0$. The
644: limit is independent of how the tetrad is extended off ${\cal I}^+$ but
645: to simplify the calculation we make the following choices adapted to
646: our conformal Bondi coordinates. We set
647: $\hat \ell^\mu =e^{2\beta}\hat \nabla^\mu u$,
648: $\hat n^\mu =\hat \nabla^\mu \ell +O(\ell)$,
649: $\hat \ell^\rho \hat \nabla_\rho \hat m^\mu=0$ and
650: $\hat \ell^\rho \hat \nabla_\rho \hat n^\mu=0$,
651: which implies
652: \begin{equation}
653: \hat n_\mu= \hat \nabla_\mu \ell
654: - \frac{\ell}{4}\hat \ell_\mu \hat\Theta +O(\ell^2).
655: \label{eq:hatn}
656: \end{equation}
657:
658: Our main calculational result is:
659: \begin{equation}
660: \hat \Psi=\frac{1}{2}\hat n^\mu \hat m^\nu \hat m^\rho \bigg(
661: \hat \nabla_\mu \hat \Sigma_{\nu\rho}
662: -\hat \nabla_\nu \hat \Sigma_{\mu\rho}\bigg)|_{\cal I^+} ,
663: \label{eq:psisigma}
664: \end{equation}
665: and that (\ref{eq:psisigma}) is independent of the freedom
666: \begin{equation}
667: \hat m^\nu \rightarrow \hat m^\nu +\lambda \hat n^\nu.
668: \label{eq:mnfreedom}
669: \end{equation}
670:
671: The result (\ref{eq:psisigma}) follows from the following sequence of
672: calculations beginning with (\ref{eq:psi}) (where evaluation at ${\cal I}^+$ is
673: assumed):
674: \begin{eqnarray}
675: -2 \hat \Psi &=&\frac{1}{\ell}
676: \hat n^\mu \hat m^\nu \hat n^\rho \hat m^\sigma
677: \hat C_{\mu\nu\rho\sigma}
678: \label{eq:stepi}\\
679: &=&\frac{1}{\ell}
680: \hat n^\mu \hat m^\nu \hat n^\rho \hat m^\sigma
681: \hat R_{\mu\nu\rho\sigma}
682: \label{eq:trace}\\
683: &=&-\frac{1}{\ell}\hat n^\mu \hat m^\nu \hat m^\rho
684: (\hat \nabla_\mu\hat \nabla_\nu \hat n_\rho
685: -\hat \nabla_\nu \hat \nabla_\mu \hat n_\rho )
686: \label{eq:commut} \\
687: &=&-\frac{1}{\ell}\hat n^\mu \hat m^\nu \hat m^\rho
688: \bigg(
689: \hat \nabla_\mu (\ell \hat \Sigma_{\nu\rho})
690: -\hat \nabla_\nu (\ell \hat \Sigma_{\mu\rho})
691: - \hat \nabla_\mu\hat \nabla_\nu (\frac{\ell\hat\Theta}{4}\ell_\rho)
692: +\hat \nabla_\nu\hat \nabla_\mu (\frac{\ell\hat\Theta}{4}\ell_\rho)
693: \bigg )
694: \label{eq:sigma}\\
695: &=&-\hat n^\mu \hat m^\nu \hat m^\rho \bigg(
696: \hat \nabla_\mu \hat \Sigma_{\nu\rho}
697: -\hat \nabla_\nu \hat \Sigma_{\mu\rho} \bigg ) \nonumber \\
698: &-&\frac{1}{\ell}\hat n^\mu \hat m^\nu \hat m^\rho \bigg(
699: \hat \Sigma_{\nu\rho} \hat \nabla_\mu \ell
700: -\hat \Sigma_{\mu\rho} \hat \nabla_\nu \ell
701: - \hat \nabla_\mu\hat \nabla_\nu (\frac{\ell\hat\Theta}{4}\ell_\rho)
702: +\hat \nabla_\nu\hat \nabla_\mu (\frac{\ell\hat\Theta}{4}\ell_\rho)
703: \bigg )
704: \label{eq:diff}\\
705: &=&-\hat n^\mu \hat m^\nu \hat m^\rho \bigg(
706: \hat \nabla_\mu \hat \Sigma_{\nu\rho}
707: -\hat \nabla_\nu \hat \Sigma_{\mu\rho} \bigg ) \nonumber \\
708: &-&\frac{1}{\ell}\hat n^\mu \hat m^\nu \hat m^\rho \bigg(
709: \hat \Sigma_{\nu\rho} \frac{\ell\hat\Theta}{4}\ell_\mu
710: -\hat \Sigma_{\mu\rho}\frac{\ell\hat\Theta}{4}\ell_\nu
711: - \hat \nabla_\mu\hat \nabla_\nu (\frac{\ell\hat\Theta}{4}\ell_\rho)
712: +\hat \nabla_\nu\hat \nabla_\mu (\frac{\ell\hat\Theta}{4}\ell_\rho)
713: \bigg )
714: \label{eq:hatn1}\\
715: &=&-\hat n^\mu \hat m^\nu \hat m^\rho \bigg(
716: \hat \nabla_\mu \hat \Sigma_{\nu\rho}
717: -\hat \nabla_\nu \hat \Sigma_{\mu\rho} \bigg ) \nonumber \\
718: &-&\frac{\hat\Theta}{4} \hat m^\nu \hat m^\rho \bigg (
719: \hat \Sigma_{\nu\rho}-n^\mu\ell^\sigma \hat R_{\mu\nu\rho\sigma}
720: \bigg )
721: \label{eq:lead}\\
722: &=&-\hat n^\mu \hat m^\nu \hat m^\rho \bigg(
723: \hat \nabla_\mu \hat \Sigma_{\nu\rho}
724: -\hat \nabla_\nu \hat \Sigma_{\mu\rho} \bigg ) \nonumber \\
725: &-&\frac{\hat\Theta}{4} \hat m^\nu \hat m^\rho \bigg (
726: \hat \Sigma_{\nu\rho}-\hat n^\mu \ell^\sigma
727: (\hat g_{\mu[\nu}\hat G_{\sigma]\rho}
728: -\hat g_{\rho[\nu}\hat G_{\sigma]\mu}) \bigg )
729: \label{eq:vac} \\
730: &=&-\hat n^\mu \hat m^\nu \hat m^\rho \bigg(
731: \hat \nabla_\mu \hat \Sigma_{\nu\rho}
732: -\hat \nabla_\nu \hat \Sigma_{\mu\rho} \bigg ) \nonumber \\
733: &-&\frac{\hat\Theta}{4} \hat m^\nu \hat m^\rho \bigg (
734: \hat\Sigma_{\nu\rho}+\frac{1}{2}\hat G_{\nu\rho})
735: \label{eq:algebra} \\
736: &=&-\hat n^\mu \hat m^\nu \hat m^\rho \bigg(
737: \hat \nabla_\mu \hat \Sigma_{\nu\rho}
738: -\hat \nabla_\nu \hat \Sigma_{\mu\rho}\bigg)
739: \label{eq:stepf}.
740: \end{eqnarray}
741: Here (\ref{eq:trace}) follows because all trace terms vanish; (\ref{eq:commut})
742: follows from the commutator of covariant derivatives; (\ref{eq:sigma}) follows
743: from (\ref{eq:Sigma}); (\ref{eq:diff}) follows from differentiation;
744: (\ref{eq:hatn1}) follows from (\ref{eq:hatn}); (\ref{eq:lead}) follows from
745: taking leading terms and using the covariant commutator; (\ref{eq:vac}) follows
746: from the vanishing of the Weyl tensor at ${\cal I}^+$; (\ref{eq:algebra})
747: follows algebraically; and (\ref{eq:stepf}) follows from (\ref{eq:einsteinsig}).
748:
749: Invariance of $\hat \Psi$ under the freedom (\ref{eq:mnfreedom}) follows from
750: noting that
751: \begin{equation}
752: \hat n^\mu \hat m^\nu \hat n^\rho \hat n^\sigma
753: \hat C_{\mu\nu\rho\sigma} =0
754: \end{equation}
755: and then following the steps analogous to (\ref{eq:stepi}) - (\ref{eq:stepf}) to
756: show
757: \begin{equation}
758: \hat n^\mu \hat m^\nu \hat n^\rho \bigg(
759: \hat \nabla_\mu \hat \Sigma_{\nu\rho}
760: -\hat \nabla_\nu \hat \Sigma_{\mu\rho}\bigg)|_{\cal I^+}=0.
761: \end{equation}
762:
763: Finally, the Weyl tensor must be scaled intrinsic to the $\tilde g_{\mu\nu}$
764: conformal frame in order to describe the radiation observed by inertial
765: observers at ${\cal I}^+$. The conformal transformation
766: $\tilde g_{\mu\nu}=\omega^2 \hat g_{\mu\nu}$ gives for the inertial radiation field
767: \begin{eqnarray}
768: \Psi:&=&-\frac{1}{2} \lim_{\Omega \rightarrow 0}\frac{1}{\Omega}
769: \tilde n^\mu {\cal Q}^\nu \tilde n^\rho {\cal Q}^\sigma
770: \tilde C_{\mu\nu\rho\sigma} \nonumber \\
771: &=&-\frac{1}{2} \omega^{-3}e^{-2i\delta} \lim_{\ell \rightarrow 0}
772: \frac{1}{\ell}
773: \hat n^\mu F^\nu \hat n^\rho F^\sigma \hat C_{\mu\nu\rho\sigma},
774: \end{eqnarray}
775: where ${\cal Q}^{\beta}=e^{-i\delta}\omega^{-1}F^\beta+\lambda {\tilde
776: n}^{\beta}$ is the same inertial polarization dyad used in describing the news
777: (\ref{eq:news}). From (\ref{eq:psisigma}), we then have
778: \begin{equation}
779: \Psi=\frac{1}{2} \omega^{-3}e^{-2i\delta}
780: \hat n^\mu F^\nu F^\rho \bigg(
781: \hat \nabla_\mu \hat \Sigma_{\nu\rho}
782: -\hat \nabla_\nu \hat \Sigma_{\mu\rho}\bigg)|_{\cal I^+}
783: \label{eq:psilim}.
784: \end{equation}
785:
786: We next need to express $\Psi$ in terms of the computational variables.
787: The straightforward way is to expand (\ref{eq:psilim}) as
788: \begin{equation}
789: \Psi=\frac{1}{2} \omega^{-3}e^{-2i\delta}
790: \hat n^\mu F^A F^B \bigg(
791: \partial_\mu \hat \Sigma_{AB}
792: -\partial_A \hat \Sigma_{\mu B}
793: - \hat \Gamma^\alpha_{\mu B}\hat \Sigma_{A \alpha}
794: +\hat \Gamma^\alpha_{A B}\hat \Sigma_{\mu\alpha}
795: \bigg)|_{\cal I^+}
796: \label{eq:psia}.
797: \end{equation}
798: and calculate the individual components of $\hat \Sigma_{\mu\nu}$ in terms of
799: those variables. This involves lengthy algebra, which is simplified by the
800: following intermediate results which hold at ${\cal I}^+$:
801: \begin{equation}
802: \hat \Sigma_{\ell\ell} =-2\partial_\ell^2 \beta
803: \end{equation}
804: \begin{equation}
805: \hat \Sigma_{\ell A} =\frac{1}{2} e^{-2H}\partial_\ell (
806: h_{AB} \partial_\ell U^B)
807: \end{equation}
808: \begin{equation}
809: \hat \Sigma_{\ell u} =\frac{1}{4} e^{2H}{\cal R}
810: +\frac{1}{4} D_A D^A e^{2H}- L^A \hat \Sigma_{\ell A}
811: \end{equation}
812: \begin{equation}
813: (\hat \nabla^\mu \ell) (\hat \nabla^\nu \ell) \hat \Sigma_{\mu\nu} =
814: \frac{1}{2} e^{-2H}(\partial_u +L^A \partial_A)(e^{-2H} D_A L^A )
815: \end{equation}
816: \begin{equation}
817: (\hat \nabla^\mu \ell) \hat \Sigma_{\mu A} =
818: \frac{1}{2} \partial_A (e^{-2H} D_B L^B )
819: \end{equation}
820: \begin{equation}
821: \hat \Sigma_{AB} = \frac{1}{2}e^{-2H} (\partial_u +{\cal L}_L )c_{AB}
822: +e^{-2H} D_A D_B e^{2H}
823: -\frac{1}{4}H_{AB}({\cal R} +3e^{-2H} D^C D_C e^{2H}).
824: \end{equation}
825: We use a Maple script to convert these expressions in terms of $\eth$
826: operators acting on the spin-weighted computational fields and construct
827: the final Fortran expression for $\Psi$.
828:
829: In inertial Bondi coordinates, (\ref{eq:psia}) reduces to
830: \begin{equation}
831: \Psi = \frac {1}{4} Q^A Q^B\partial_u^2 c_{AB} = \partial_u^2
832: \partial_l J|_{{\cal I}}^+ .
833: \end{equation}
834: This is related to the expression for the news function in
835: inertial Bondi coordinates by
836: \begin{equation}
837: \Psi =\partial_u N.
838: \label{eq:PsiNu}
839: \end{equation}
840: However, as in the case of the news, the full expression (\ref{eq:psia}) for
841: $\Psi$ must be used in the code. This introduces additional challenges to
842: numerical accuracy due to the large number of terms and the appearance of third
843: angular derivatives.
844:
845: \subsection{Linearized expressions}
846:
847: The general nonlinear representation of $\Psi$ in (\ref{eq:psia}) in terms of
848: the computational variables is quite long but reduces to a simpler form in the
849: linearized approximation, i.e. to first order in perturbations off a Minkowski
850: background. In terms of the spin-weighted fields $J=h_{AB}q^A Q^B /2$
851: and $L=L^A q_A$, we find
852: \begin{equation}
853: \Psi=-\frac{1}{2}\hat\Sigma_{\ell u}\eth L +\partial_u \hat\Sigma_J
854: -\frac{1}{2}\eth \hat\Sigma_u
855: -\frac{1}{2} \partial_u J\left(\hat\Sigma_{\ell u}+\hat\Sigma_{K}\right)
856: \label{e-Psil1}
857: \end{equation}
858: (evaluated at ${\cal I}^+$), where the only nonvanishing zeroth order parts of
859: $\hat\Sigma_{\mu\nu}$ are
860: \begin{equation}
861: \hat\Sigma_K \equiv \frac{1}{2} q^A \bar{q}^B \hat\Sigma_{AB}
862: =-\frac{1}{2},\;\;\hat\Sigma_{\ell u}=\frac{1}{2}
863: \end{equation}
864: and the required first order components are
865: \begin{equation}
866: \hat\Sigma_J \equiv \frac{1}{2} q^A q^B \hat\Sigma_{AB}
867: =\eth^2 H -\frac{1}{2}J
868: +\frac{1}{2} \partial_u \partial_\ell J
869: \end{equation}
870: \begin{equation}
871: \hat\Sigma_u \equiv q^A \hat\Sigma_{uA}
872: =\frac{1}{4}\eth^2 \bar{L} +\frac{1}{4}\eth\bar{\eth}L +\frac{1}{2} L.
873: \end{equation}
874: Then (\ref{e-Psil1}) reduces to
875: \begin{equation}
876: \Psi=\frac{1}{2}\partial_u^2 \partial_\ell J -\frac{1}{2}\partial_u J
877: -\frac{1}{2}\eth L -\frac{1}{8} \eth^2( \eth \bar L +\bar \eth L)
878: + \partial_u \eth^2 H.
879: \label{eq:linPsi}
880: \end{equation}
881:
882: In the same approximation, the news function is given by
883: \begin{equation}
884: N=\frac{1}{4}q^A q^B \bigg (\partial_u c_{AB}
885: +2 D_A D_B ( \omega +2H) \bigg ) =\frac{1}{2} \partial_u \partial_\ell J
886: +\frac{1}{2} \eth^2(\omega +2H).
887: \label{eq:linN}
888: \end{equation}
889: Using the asymptotic relations
890: \begin{eqnarray}
891: \partial_u J&=&- \eth L
892: \label{eq:linJu} \\
893: \partial_u \omega &=&-\frac{1}{4}(\eth \bar L + \bar \eth L) ,
894: \label{eq:linomegau}
895: \end{eqnarray}
896: which arise from the linearized versions of (\ref{eq:asymHu}) and
897: (\ref{eq:omegadot}), it is easy to see that (\ref{eq:PsiNu}), i.e. $\Psi
898: =\partial_u N$, still holds in the linearized approximation. (In the nonlinear
899: case, the derivative along the generators of ${\cal I}^+$ is $\hat n^\mu
900: \partial_\mu =e^{-2H}(\partial_u +L^A \partial_A)$ and (\ref{eq:PsiNu}) must be
901: modified accordingly.)
902:
903: The linearized expressions (\ref{eq:linPsi}) and (\ref{eq:linN}) provide a
904: starting point to compare the advantages between computing the radiation via the
905: Weyl component $\Psi$ or the news function $N$. The troublesome terms
906: involve $L$, $H$ and $\omega$, which all vanish in inertial Bondi coordinates.
907: One main difference is that $\Psi$ contains third order angular
908: derivatives, e.g. $\eth^3 \bar L$, as opposed to second angular derivatives for
909: $N$. This means that the smoothness of the numerical error is more crucial in
910: the $\Psi$ approach. Balancing this, another main difference is that $N$
911: contains the $\eth^2 \omega$ term, which is a potential source of numerical
912: error since $\omega$ must be propagated across patch boundaries via
913: (\ref{eq:omegadot}).
914:
915:
916: \subsection{Summary of the gravitational radiation calculation}
917: \label{sec:sum}
918:
919: The characteristic Einstein equations are evolved in a domain between an inner
920: radial boundary at the interior worldtube, and an outer boundary at future null
921: infinity. Initial data for $J(u,r,x^A)$ is required at $u=0$. This data is
922: constraint-free so that, in the absence of an exact solution or other
923: prescription of the data, we can simply set $J(0,r,x^A)=0$. Alternatively, in
924: order to reduce spurious initial radiation, we can set the Newman-Penrose Weyl
925: tensor component $\Psi_0(0,r,x^A) =0$, which determines $J(0,r,x^A)=0$ when
926: continuity conditions are imposed at the inner worldtube. The metric data from a
927: Cauchy evolution is interpolated onto the inner worldtube to extract the
928: boundary data for the characteristic evolution. This extraction process involves
929: carrying out the complicated Jacobian transformation between the Cartesian
930: coordinates used in the Cauchy evolution and the spherical null coordinates
931: used in the characteristic evolution. The full details are given in~\cite{ccm}.
932: The result is boundary data for $J,\beta,U,Q,V$ on the worldtube, which supply
933: the integration constants for a radial numerical integration of (\ref{eq:beta}),
934: (\ref{eq:wua}), (\ref{eq:wq}) and (\ref{eq:ww}), in that order. Given the
935: initial data $J(0,r,x^A)$, this leads to complete knowledge of the metric on the
936: initial null cone. Then (\ref{eq:wev}) gives an expression for $J_{,ur}$, which
937: is used to determine $J$ on the ``next'' null cone, so that the process can be
938: repeated to yield the complete metric throughout the domain, which extends to
939: ${\cal I}^+$.
940:
941: Before the gravitational radiation is calculated from the metric in the
942: neighborhood of ${\cal I}^+$, it is necessary to compute the auxiliary
943: variables $\omega(u,x^A)$ and $\delta(u,x^A)$ which determine the inertial
944: polarization dyad in which to measure the news function $N$ or Weyl component
945: $\Psi$. Given a solution of (\ref{eq:conf}) for the initial value of
946: $\omega(0,x^A)$, its evolution is computed by integrating (\ref{eq:omegadot}).
947: (If $J=0$ initially, then $\omega=1$ is the solution to (\ref{eq:conf}).
948: Otherwise, $\omega$ is initiated by solving a 2-dimensional elliptic equation.)
949: Similarly, fixing the initial polarization basis by $\delta(0,x^A) =0$, its
950: evolution is computed by integrating (\ref{eq:evphase}). Then the news $N$ is
951: given by (\ref{eq:news}) (or in spin-weighted form by the formulas in Appendix
952: B of Ref.~\cite{highp}) and $\Psi$ is given by (\ref{eq:psia}).
953:
954: The above procedure computes $N$ or $\Psi$ as functions of the code coordinates
955: $(u,x^A)$, rather than inertial coordinates. In the linearized case, which is
956: used for the tests in Sec.~\ref{sec:extraction}, the change to inertial
957: coordinates is a second-order effect that can be neglected. However, that is not
958: the case in general and the full procedure is described in Sec. IV B of
959: Ref.~\cite{highp}.
960:
961: \section{Patching the sphere}
962: \label{sec:patches}
963:
964: The nonsingular description of smooth tensor fields on the sphere requires more
965: than a single coordinate patch. Here we consider the stereographic treatment
966: which uses 2 coordinate patches, and the cubed-sphere treatment, which uses 6
967: patches. In both cases the metric $q_{AB}$ of the unit sphere is expressed in
968: terms of a complex dyad $q_A$ (satisfying $q^Aq_A=0$, $q^A\bar q_A=2$,
969: $q^A=q^{AB}q_B$, with $q^{AB}q_{BC}=\delta^A_C$ and $ q_{AB}
970: =\frac{1}{2}\left(q_A \bar q_B+\bar q_A q_B\right)$). The dyads for each patch
971: are related by spin transformations at points common
972: to more than one patch.
973:
974: \subsection{Circular stereographic patches}
975:
976: In stereographic coordinates, the sphere is covered with
977: two patches, one for each hemisphere.
978: The North hemisphere is covered by the complex stereographic coordinate
979: $\xi_N=\eta_N+i\rho_N$, which is related to
980: standard $(\theta, \phi)$ angular coordinates by
981: $\xi_N=\tan(\theta/2)e^{i\phi}$ and which
982: is regular on the entire sphere except for the South pole.
983: The South hemisphere is covered by the complex stereographic coordinate
984: $\xi_S=1/\xi_N=\eta_S+i\rho_S=\cot(\theta/2)e^{-i\phi}$,
985: which is singular at the North pole.
986: Every point on the sphere is covered by at least one of the patches, and there
987: is a region around the equator where points are covered by both patches.
988: In this overlap region between the two patches, a scalar function $F$ with
989: value $F_N(\xi_N)$ on the North patch has the value
990: $F_S(\xi_S=1/\xi_N)$ on the South patch. For a function $F$ of spin-weight
991: $s$, $F_S(\xi_S=1/\xi_N)=F_N(\xi_N)(-1)^se^{-2is\Phi}$.
992:
993: In the $x^A=(\eta,\rho)$ coordinates, the unit sphere
994: metric in each patch is given by
995: \begin{equation}
996: q_{AB} dx^A dx^B = \frac{4}{P^2}(d\eta^2+d\rho^2),
997: \end{equation}
998: where
999: \begin{equation}
1000: P=1+\eta^2+\rho^2.
1001: \end{equation}
1002: The equator corresponds to the circle
1003: \begin{equation}
1004: \sqrt{\eta^2+\rho^2} = 1.
1005: \end{equation} We fix
1006: the dyad by the explicit choice
1007: \begin{equation}
1008: q^A=\frac{P}{2}(1,i)\, , \quad i=\sqrt{-1}.
1009: \end{equation}
1010:
1011: In the composite mesh method, all boundary points of one patch are interior
1012: points of another patch. The overlapping of the patches is key to the
1013: stability of this method. The two stereographic
1014: coordinate patches must both extend
1015: beyond the equator. In the scheme originally used to implement the computational
1016: $\eth$-formalism~\cite{eth} in the characteristic code, the North and South
1017: patches were represented by square $(\eta,\rho)$ grids. In the scheme implemented for
1018: meteorological studies~\cite{browning}, circular masks are applied so that the
1019: computational grids extend only a few zones beyond the equator. Here we adopt
1020: this circular grid boundary but place it a fixed geometrical distance past the
1021: equator, i.e. the grid boundary for the North patch is a circle lying in the
1022: South patch. The finite buffer zone between the equator and the grid boundary
1023: allows for angular dissipation, as developed in Sec.~\ref{sec:angdiss}, to damp
1024: the high frequency intergrid interpolation error before it crosses the equator.
1025: This protects measurements of the news function (or $\Psi$) in the North patch,
1026: which involve two (or three) angular derivatives, from substantial
1027: contamination by the interpolation error at the patch boundary.
1028:
1029: We discretize the stereographic coordinates according to
1030: \begin{equation}
1031: \eta_i = -1 + (i-O-1)\Delta \, , \quad
1032: \rho_j = -1 + (j-O-1)\Delta
1033: \label{eq:sqgrid}
1034: \end{equation}
1035: where, following the notation in~\cite{browning}, $O$ is the number of points
1036: (overlapping points) by which each grid extends beyond the equator
1037: and the indices range over $1 \le i,j \le M+1+2O$, with $M^2$ being the number of
1038: grid points inside the equator. The grid spacing $\Delta$ depends on $M$ according to
1039: \begin{equation}
1040: \Delta = \frac{2}{M}.
1041: \end{equation}
1042: The square grid determined by (\ref{eq:sqgrid}) ranges over
1043: \begin{equation}
1044: (\eta_i,\rho_j) \in \left( -1-O\Delta, 1+O\Delta \right)
1045: \label{eq:qpgrid}
1046: \end{equation}
1047: in each patch.
1048:
1049: In the original square patch method~\cite{eth}, the evolution algorithm is
1050: applied to the entire set of points in the square $(\eta,\rho)$ grid, with the
1051: field values at the resulting ghost points supplied by interpolation from the
1052: other patch. In the circular patch method~\cite{browning}, the evolution
1053: algorithm is only applied to points inside a circle $r=\sqrt{\eta^2+\rho^2}$,
1054: where $r>1$ so that the boundary lies a small distance past the equator. In
1055: convergence tests, the number of overlap points determined by $O$ is adjusted so
1056: that $r$ is at a fixed position for all grids, i.e. $O$ scales as $1/\Delta$.
1057: The grid points outside this circle are either ghost points or inactive. The
1058: circular patch is clearly more economical than using a square patch and avoids
1059: the error introduced by the large stereographic grid stretching near the corners
1060: of the square.
1061:
1062: When the finite-difference stencil is used near the boundary of the active grid
1063: points, field values required at the ghost points outside a circular patch are
1064: interpolated from values at interior points of the opposite patch. The
1065: algorithm for determining the value of a scalar function $F_N$ at a ghost point
1066: in the North patch starts with the determination of the ghost point's
1067: coordinates in the overlapping South patch, followed by the interpolation of the
1068: value of the function $F_S$ at the ghost point, i.e. the $F_N$ ghost point
1069: values are obtained by interpolation via the $F_S$ active grid values.
1070:
1071: Let $R_E$ be the width of the finite-difference stencil divided by $2 \Delta$.
1072: In the circular patch method, we define the {\em active} finite difference
1073: grid, i.e. the grid points to which the
1074: evolution algorithm is applied, by
1075: \begin{equation}
1076: \sqrt{\eta_i^2+\rho_j^2} \le 1 + (O-R_E)\Delta,
1077: \label{eq:evmask}
1078: \end{equation}
1079: where $O > R_E$. Stability of the composite mesh method requires that the
1080: interpolation stencil for the ghost points for one patch lies below the equator
1081: in the other patch. Those requirements give a minimum value of $O$ but a larger
1082: value may be necessary to establish an effective buffer zone for the dissipation
1083: to attenuate the interpolation error before it enters the opposite patch. The
1084: optimum value of $O$ in order to avoid instability or inaccuracy, needs to be
1085: established by experiment. (Too large a value would lead to inaccuracy due to
1086: the stretching of the stereographic grid.)
1087:
1088: \subsection {The cubed sphere}
1089:
1090: In the cubed-sphere approach, developed for meteorological studies
1091: in~\cite{ronchi} and later for numerical relativity in~\cite{Thornburgah}, 6
1092: coordinate patches on the sphere are obtained by projecting the 6 faces of a
1093: circumscribed cube. The method has recently been applied to characteristic
1094: evolution in~\cite{reisswig} and independently in~\cite{roberto}. Here we follow
1095: the notation of~\cite{reisswig}, except we denote the angular coordinates by
1096: $(\phi_1,\phi_2)$ (rather than by $(\rho,\sigma)$). In addition, in order to
1097: ensure that the coordinates and dyads on each patch are consistently
1098: right-handed, with the vector cross-product vector pointing out of the sphere,
1099: we introduce some sign changes in the conventions used in~\cite{reisswig} for
1100: the coordinate transformations between the patches and in the definition of the
1101: dyad $q^A$. These conventions simplify the interpatch transformation of
1102: spin-weighted quantities.
1103:
1104: Given Cartesian coordinates $(x,y,z)$, we define
1105: angular coordinates $x^A=(\phi_1,\phi_2)$ on the 2-sphere $x^2+y^2+z^2=1$
1106: by means of the six patches $(x_\pm,y_\pm,z_\pm)$, where
1107: \begin{eqnarray}
1108: x_\pm&:&\;
1109: \phi_1=\arctan \left(\pm \frac{y}{x}\right),
1110: \phi_2=\arctan \left( \frac{z}{x}\right)
1111: \nonumber \\
1112: y_\pm&:&\;
1113: \phi_1=\arctan \left( \pm\frac{z}{y}\right),
1114: \phi_2=\arctan \left( \frac{x}{y}\right) \nonumber \\
1115: z_\pm&:&\;
1116: \phi_1=\arctan \left( \pm \frac{x}{z}\right),
1117: \phi_2=\arctan \left( \frac{y}{z}\right).
1118: \label{e-six2C}
1119: \end{eqnarray}
1120:
1121: In each patch, the range of the coordinates is
1122: $-\pi/4 \le\phi_1,\phi_2\le\pi/4$ and the metric is
1123: \begin{equation}
1124: ds^2=\left(1- \sin^2 \phi_1 \; \sin^2 \phi_2 \right)^{-2}
1125: \bigg(
1126: \cos^2\phi_2 \; d\phi_1^2 + \cos^2\phi_1 \; d\phi_2^2
1127: - \frac{1}{2} \sin(2\phi_1) \sin(2\phi_2) \; d\phi_1 \, d\phi_2
1128: \bigg).
1129: \label{e-m1}
1130: \end{equation}
1131: As a simple dyad representing (\ref{e-m1}), we choose
1132: \begin{equation}
1133: q_A=\bigg(
1134: \frac{(\theta_c-i \theta_s)\cos\phi_2}{4\theta_c^2 \theta_s^2},
1135: \frac{(\theta_c+i \theta_s)\cos\phi_1}{4\theta_c^2 \theta_s^2}
1136: \bigg),\;\;
1137: q^A=\bigg(
1138: 2\theta_c\theta_s\frac{\theta_s
1139: -i \theta_c}{\cos\phi_2},
1140: 2\theta_c\theta_s\frac{\theta_s
1141: +i \theta_c}{\cos\phi_1}
1142: \bigg),
1143: \label{e-dyad}
1144: \end{equation}
1145: where
1146: \begin{equation}
1147: \theta_c=\sqrt{\frac{1-\sin\phi_1\sin\phi_2}{2}},
1148: \theta_s=\sqrt{\frac{1+\sin\phi_1\sin\phi_2}{2}}.
1149: \end{equation}
1150: The operator $\eth$ acting on
1151: a field $S$ with spin-weight $s$ is
1152: $\eth S=q^A \partial_A S + s \Gamma S$ where, with the present conventions,
1153: \begin{eqnarray}
1154: \Gamma&=&
1155: \frac{\cos^2\phi_1\cos^2\phi_2(\sin\phi_1+\sin\phi_2)
1156: +(\cos^2\phi_1-\cos^2\phi_2)(\sin\phi_2-\sin\phi_1)}
1157: {4 \theta_c \cos\phi_2 \cos\phi_1} \nonumber \\
1158: &-&i\frac{\cos^2\phi_1 \cos^2\phi_2 (\sin\phi_1-\sin\phi_2)
1159: +(\cos^2\phi_2-\cos^2\phi_1)(\sin\phi_1+\sin\phi_2)}
1160: {4 \theta_s \cos\phi_2 \cos\phi_1}.
1161: \label{e-Gamma}
1162: \end{eqnarray}
1163:
1164: We introduce ghost zones in the usual manner along the boundaries of each patch,
1165: and couple the patches together by interpolating the field variables from
1166: neighboring patches to each ghost point. With the definition (\ref{e-six2C}),
1167: the angular coordinate~$\phi_1$ or~$\phi_2$ perpendicular to an interpatch
1168: boundary is always common to both adjacent patches. This greatly simplifies
1169: interpatch interpolation, since it only needs to be done in
1170: 1~dimension, parallel to the boundary.
1171:
1172: \section{Comparison between stereographic and cubed-sphere methods}
1173: \label{sec:stests}
1174:
1175: We carry out a test of wave propagation on the sphere to compare the accuracy of
1176: using circular stereographic patches with the cubed-sphere methods, with
1177: emphasis on the accuracy of the angular derivatives required in waveform
1178: extraction by the news and $\Psi_4$ approaches. The test allows direct
1179: comparison between the stereographic and cubed-sphere treatments without
1180: introducing the complications of characteristic evolution and the explicit
1181: calculations of the news or $\Psi$.
1182:
1183: The test is based upon solutions to the 2D wave equation
1184: \begin{equation}
1185: - \partial_t ^2 \Phi + \eth \bar \eth \Phi = 0,
1186: \end{equation}
1187: where $\Phi =cos(\omega t) Y_{\ell m}$, $\omega =\sqrt{\ell(\ell+1)}$
1188: and $Y_{\ell m}$ are spherical harmonics.
1189:
1190: For the case $\ell=m=2$, we compare test results for the stereographic
1191: grid with circular patches and the
1192: cubed-sphere grid. For the stereographic grid, the simulations are run with
1193: $M^2$ grid points in each patch, for $M=100$ and $M=120$. The
1194: corresponding cubed sphere runs keep the number of grid-cells covering the
1195: sphere the same as for the stereographic case, not counting those cells that
1196: overlap with another patch. For $M^2$ stereographic grid points there are
1197: $\approx \pi M^2 / 4$ grid-cells inside the equator on each hemisphere. In the
1198: cubed sphere grid, with $N^2$ points per patch, the entire sphere is covered by
1199: $6 \times N^2$ points. Equating the number of cells for the entire sphere gives
1200: $N^2 \approx (\pi/12) M^2$. The above values of $M$ then
1201: correspond to $N = 51, 61$.
1202: The tests are run until $t=120$.
1203:
1204: Angular dissipation is necessary for the stability of the stereographic runs.
1205: For grid size $\Delta$, it was added in the finite-difference form
1206: \begin{equation}
1207: \partial_t^2 \Phi \rightarrow \partial_t^2 \Phi
1208: + \epsilon \Delta^3 {\cal D}^4 \partial_t\Phi ,
1209: \end{equation}
1210: where ${\cal D}^4 \Phi =\left (\frac{P^2}{4}({\cal D}_{+\eta}{\cal D}_{-\eta}
1211: +{\cal D}_{+\rho}{\cal D}_{-\rho})\right )^2\Phi $, where ${\cal D}_{+}$ (or
1212: ${\cal D}_{-}$) indicates the forward (or backward) difference operator in the
1213: indicated direction. Experimentation with tuning the dissipation revealed that a
1214: small value $\epsilon = 0.01$ is sufficient to suppress high frequency error.
1215: The finite difference stencil (taking dissipation into account) has width
1216: $R_E=2$. Using a $4th$ order Lagrange interpolator and by tuning the number $O$
1217: of overlapping points, we obtained good results with $O=5$. Angular dissipation
1218: was not used in the cubed sphere runs.
1219:
1220: We use the $L_\infty$ norm to measure the error
1221: \begin{equation}
1222: {\cal E}(\Phi) =||\Phi_{numeric}-\Phi_{analytic}||_\infty.
1223: \end{equation}
1224: We measure the convergence rate for ${\cal E}(\Phi)$ at a given time $t$,
1225: for two grid sizes $\Delta_1$ and $\Delta_2$, by
1226: \begin{equation}
1227: {\cal R} = \frac{\log_2 \big ({\cal E}(\Phi)_{\Delta_1}
1228: / {\cal E}(\Phi)_{\Delta_2} \big )}
1229: {\log_2 \big (\Delta_2 / \Delta_1 \big )}.
1230: \end{equation}
1231: Convergence rates for other quantities are measured analogously. For a given
1232: grid, we measure the error for the circular patches in the North
1233: hemisphere; while for the cubed-sphere method we measure the error on the
1234: $(+x,+y,+z)$ patch, excluding ghost points at the edges of the patch. The
1235: finite difference approximations for the codes are designed to be second
1236: order accurate.
1237:
1238: Excellent second-order convergence of ${\cal E}(\Phi)$, based upon the
1239: $M=100$ and $M=120$ grids, is evident for both methods from the results listed in
1240: Table~\ref{tab:ConvPhi}. The time dependence of the error plots in
1241: Fig.~\ref{fig:Phin_error}, shows that the cubed-sphere error is
1242: $\approx \frac{1}{3}$ the stereographic error.
1243:
1244: \begin{table}[htp]
1245: \begin{center}
1246: \begin{tabular}[c]{|c|c|c|c|c|}
1247: \hline
1248: ALGORITHM &
1249: t=1.2 &
1250: t=12 &
1251: t = 102 &
1252: t = 120
1253:
1254: \\ \hline \hline
1255: %%%%%
1256: circular patch & $2.002$ & $1.988$ & $1.994$ & $1.999$
1257: \\ \hline
1258: %%%%%
1259: %%%%%
1260: cubed-sphere & $1.994$ & $1.970$ & $1.982$ & $1.985$
1261: \\ \hline
1262: %%%%%
1263: \end{tabular}
1264: \caption{Convergence rates for ${\cal E}(\Phi)$, obtained with
1265: the $L_\infty$ norm using the $M=100$ and $M=120$ grids.}
1266: \label{tab:ConvPhi}
1267: \end{center}
1268: \end{table}
1269:
1270:
1271: \begin{figure}[htp] % figure placement: here, top, bottom, or page
1272: \centering
1273: \psfrag{time}{t}
1274: \psfrag{circular}{circular patch}
1275: \psfrag{sixpatch}{cubed-sphere}
1276: \psfrag{error}[c][c]{${\cal E}(\Phi)$ }
1277: \includegraphics*[width=10cm]{Figure01}
1278: \caption{Comparison of the $L_\infty$ error ${\cal E}(\Phi)$ vs $t$,
1279: for the highest resolution runs using the circular patch method
1280: and the cubed-sphere method. }
1281: \label{fig:Phin_error}
1282: \end{figure}
1283:
1284:
1285: A more important test to assess the error relevant to gravitational wave
1286: extraction is to measure the error in $\eth^2\Phi$, since second angular
1287: derivatives enter in the computation of the Bondi news. The convergence rates,
1288: measured with the $L_\infty$ norm, are shown in Table~\ref{tab:Conveth2Phi}. The
1289: circular patch and cubed-sphere results indicate clean second order convergence
1290: up to the final run time at $t=120$. The plots of the error versus time in
1291: Fig.~\ref{fig:eth2Phin_error} show that the error for the cubed-sphere is about
1292: $\frac{2}{3}$rd the stereographic error.
1293:
1294: \begin{table}[htp]
1295: \begin{center}
1296: \begin{tabular}[c]{|c|c|c|c|c|}
1297: \hline
1298: ALGORITHM &
1299: t=1.2 &
1300: t=12 &
1301: t = 102 &
1302: t = 120
1303:
1304: \\ \hline \hline
1305: %%%%%
1306: circular patch & $2.022$ & $1.945$ & $1.992$ & $2.006$
1307: \\ \hline
1308: %%%%%
1309: %%%%%
1310: cubed-sphere & $1.954$ & $2.019$ & $1.997$ & $1.971$
1311: \\ \hline
1312: %%%%%
1313: \end{tabular}
1314: \caption{Convergence rates for the $L_\infty$ error ${\cal E}(\eth^2\Phi)$
1315: for various times $t$, obtained using the two highest resolution runs.
1316: }
1317: \label{tab:Conveth2Phi}
1318: \end{center}
1319: \end{table}
1320:
1321:
1322: \begin{figure}[htp] % figure placement: here, top, bottom, or page
1323: \centering
1324: \psfrag{time}{t}
1325: \psfrag{circular}{circular patch}
1326: \psfrag{sixpatch}{cubed-sphere}
1327: \psfrag{error}[c][c]{${\cal E}(\eth^2 \Phi)$ }
1328: \includegraphics*[width=10cm]{Figure02}
1329: \caption{The $L_\infty$ error ${\cal E}(\eth^2 \Phi)$ vs
1330: $t$ for the highest resolution runs is compared using
1331: the circular patch method and the cubed-sphere method. }
1332: \label{fig:eth2Phin_error}
1333: \end{figure}
1334:
1335:
1336: Similarly, it is important for the purpose of gravitational wave extraction
1337: using the Weyl tensor to measure the error in $\bar \eth \eth^2\Phi$, since third angular
1338: derivatives enter into the
1339: computation of $\Psi$. The corresponding convergence rates are shown in
1340: Table~\ref{tab:Convethbeth2Phi}. Now the cubed-sphere method shows poor
1341: convergence of the $L_\infty$ error at early times. The underlying error is
1342: generated at the corners where three patches meet, as indicated by the improved
1343: convergence rate measured using the $L_2$ norm. Apparently some built-in dissipation of the
1344: evolution algorithm smooths this patch-boundary error and second order
1345: convergence is evident by by $t=102$. To a much smaller extent,
1346: the $L_\infty$ error for the circular
1347: patch method also shows some deviation from second order convergence at early
1348: times, but clean second order convergence is evident by
1349: $t=12$.
1350:
1351: The magnitude of the $L_\infty$ error in $\bar \eth \eth^2\Phi$ is plotted vs
1352: time in Fig.~\ref{fig:ethbeth2Phin_error}. Until about $t=60$,
1353: the cubed-sphere method has the largest error. But at the end of the run
1354: at $t=120$ the cubed-sphere error is about $\frac{4}{5}$ the
1355: stereographic error.
1356:
1357: Surface plots of the error at the final run time are shown in
1358: Fig.~\ref{fig:ethbeth2Phin_2Derror}. The circular patch and cubed-sphere errors
1359: are both smooth, as would be expected of the second order truncation error
1360: arising from the finite differencing. For the circular patch, this shows that
1361: dissipation in the buffer zone surrounding the equator effectively guards
1362: against the high frequency error introduced at the patch boundary. Our results
1363: for the stereographic method justify its use in the comparison of the news and
1364: Weyl tensor extraction strategies in Sec.~\ref{sec:extraction}.
1365:
1366: \begin{table}[htp]
1367: \begin{center}
1368: \begin{tabular}[c]{|c|c|c|c|c|}
1369: \hline
1370: ALGORITHM &
1371: t=1.2 &
1372: t=12 &
1373: t = 102 &
1374: t = 120
1375:
1376: \\ \hline \hline
1377: %%%%%
1378: circular patch, $L_\infty$ norm & $2.278$ & $2.032$ & $1.988$ & $2.009$
1379: \\ \hline
1380: %%%%%
1381: %%%%%
1382: cubed-sphere, $L_\infty$ norm & $1.108$ & $0.882$ & $2.009$ & $1.959$
1383: \\ \hline
1384: cubed-sphere, $L_2$ norm & $1.883$ & $1.983$ & $1.981$ & $1.959$
1385: \\ \hline
1386:
1387: %%%%%
1388: \end{tabular}
1389: \caption{Convergence rates for the error ${\cal E}(\bar \eth
1390: \eth^2\Phi)$. For the cubed-sphere method the dominant error arises
1391: at the patch corners, which is revealed by the comparison of the
1392: $L_\infty$ and $L_2$ errors. This can be understood in terms of the
1393: the inter-patch interpolation stencil which is partially
1394: off-centered near the corners, where the error is greatest.
1395: The inherent numerical dissipation of the evolution algorithm
1396: keeps this localized, non-smooth noise from growing,
1397: while smoother error from other regions grows linearly in time
1398: (see Fig.~\ref{fig:ethbeth2Phin_error}). The net
1399: effect is that at late times both the $L_\infty$ and the
1400: $L_2$ norms of the error in the third derivative
1401: show second order convergence, while early in the evolution
1402: the $L_\infty$ shows only first order convergence.
1403: }
1404: \label{tab:Convethbeth2Phi}
1405: \end{center}
1406: \end{table}
1407:
1408:
1409:
1410: \begin{figure}[htp] % figure placement: here, top, bottom, or page
1411: \centering
1412: \psfrag{time}{t}
1413: \psfrag{circular}{circular patch}
1414: \psfrag{sixpatch}{cubed-sphere}
1415: \psfrag{error}[c][c]{${\cal E}(\bar \eth \eth^2 \Phi)$ }
1416: \includegraphics*[width=10cm]{Figure03}
1417: \caption{The error ${\cal E}(\bar \eth \eth^2
1418: \Phi)$ vs $t$ is compared for the highest resolution runs using
1419: the circular patch method and the cubed-sphere method.
1420: }
1421: \label{fig:ethbeth2Phin_error}
1422: \end{figure}
1423:
1424:
1425: \begin{figure}[htp] % figure placement: here, top, bottom, or page
1426: \centering
1427: \psfrag{xmtickgz}[c][b]{$-\pi/4$}
1428: \psfrag{xztickgz}[c][b]{$0$}
1429: \psfrag{xptickgz}[c][l]{$+\pi/4$}
1430: \psfrag{ymtickgz}[c][r]{$-\pi/4$}
1431: \psfrag{yztickgz}[c][r]{$0$}
1432: \psfrag{yptickgz}[c][r]{$+\pi/4$}
1433:
1434: \psfrag{xmtick}[c][b]{$-1$}
1435: \psfrag{xztick}[c][b]{$0$}
1436: \psfrag{xptick}[c][b]{$+1$}
1437: \psfrag{ymtick}[c][b]{$-1$}
1438: \psfrag{yztick}[c][b]{$0$}
1439: \psfrag{yptick}[c][b]{$+1$}
1440:
1441: \includegraphics*[width=8.5cm]{Figure04}
1442: \includegraphics*[width=8.5cm]{Figure05}
1443: \caption{2D snapshots of the error in
1444: $\bar \eth \eth^2\Phi$ at $t=120$ for the
1445: on the North hemisphere for the circular patch method (left plot),
1446: and for a cubed-sphere patch (right plot).
1447: For the sake of plot clarity these
1448: 2D snapshots use only every third data-point
1449: along each axis. The third angular derivatives are smooth for
1450: both methods
1451: }
1452: \label{fig:ethbeth2Phin_2Derror}
1453: \end{figure}
1454: \clearpage
1455:
1456: \section{Comparison of news and Weyl tensor extraction}
1457: \label{sec:extraction}
1458:
1459: Here we compare the accuracy of waveform extraction by computing the news
1460: function $N$ or the Weyl tensor component $\Psi$ in a linearized gravitational
1461: wave test problem. The computations are carried out by the procedure described
1462: in Sec.~\ref{sec:sum}. In accord with (\ref{eq:PsiNu}), the $\Psi$ computation
1463: yields an alternative numerical value for the news
1464: \begin{equation}
1465: N_\Psi = N|_{u=0} +\int_0^u \Psi du,
1466: \label{eq:npsi}
1467: \end{equation}
1468: where $N=N_\Psi$ in the analytic problem.
1469: We compare the two extraction methods in terms of the errors
1470: in $N$ and $N_\Psi$ obtained using the stereographic method.
1471:
1472: We base the test on a class of solutions in Bondi-Sachs form to the linearized
1473: vacuum Einstein equation on a Minkowski background given in Sec. 4.3
1474: of~\cite{BS-lin}. The solution allows us to make convergence checks of the
1475: Bondi-Sachs metric quantities as well as the news function. The solutions are
1476: expressed in terms of spin-weighted spherical harmonics ${}_s Y_{\ell
1477: m}$~\cite{newp,golm}, modified to avoid mixing of the $m$ and $-m$ components
1478: when extracting the real part according to~\cite{mod}
1479: \begin{equation}
1480: {}_s R_{\ell m} = \frac{1}{\sqrt{2}} \left[{}_s Y_{\ell m}
1481: +(-1)^m {}_s Y_{\ell -m}\right] \mbox{ for } m>0,\;
1482: {}_s R_{\ell m} = \frac{i}{\sqrt{2}} \left[(-1)^m{}_s Y_{\ell m}
1483: -{}_s Y_{\ell -m} \right]\mbox{ for } m<0 ,\;
1484: {}_s R_{\ell 0} = {}_s Y_{\ell 0}.
1485: \end{equation}
1486: Ref.~\cite{mod} gives explicit expressions for the ${}_s R_{\ell m}$
1487: in stereographic coordinates.
1488:
1489: Following~\cite{reisswig}, these
1490: linearized solutions have Bondi-Sachs variables
1491: \begin{eqnarray}
1492: J&=& \sqrt{(\ell -1)\ell(\ell+1)(\ell+2)}\;
1493: {}_2 R_{\ell m} \Re(J_\ell(r) e^{i\nu u}), \;
1494: U= \sqrt{\ell(\ell+1)}\;{}_1R_{\ell m} \Re(U_\ell(r) e^{i\nu u}),
1495: \nonumber \\
1496: \beta&=& R_{\ell m} \Re(\beta_\ell e^{i\nu u}),
1497: \; W_c= R_{\ell m} \Re(W_{c\ell}(r) e^{i\nu u}),
1498: \label{e-an}
1499: \end{eqnarray}
1500: where $W_c$ determines the perturbation in $V$.
1501: Here $J_\ell(r)$, $U_\ell(r)$, $\beta_\ell$, $W_{c\ell}(r)$ are in general
1502: complex, and taking the real part leads to $\cos(\nu u)$ and $\sin(\nu u)$
1503: terms. The quantities $\beta$ and $W_c$ are real, while $J$ and $U$ are
1504: complex. We
1505: require a solution that is well-behaved at future null infinity, and is
1506: well-defined for $r \ge r_0>0$, where $r_0$ is the inner boundary.
1507: We find in the case $\ell=2$
1508: \begin{eqnarray}
1509: \beta_2&=&\beta_0 \nonumber \\
1510: J_2(r)&=&\frac{24\beta_0 +3 i \nu C_1 - i \nu^3 C_2}{36}+\frac{C_1}{4 r}
1511: -\frac{C_2}{12 r^3} \nonumber \\
1512: U_2(r)&=&\frac{-24i\nu \beta_0 +3 \nu^2 C_1 - \nu^4 C_2}{36} +\frac{2\beta_0}{r}
1513: +\frac{C_1}{2 r^2} +\frac{i\nu C_2}{3 r^3} +\frac{C_2}{4 r^4} \nonumber \\
1514: W_{c2}(r)&=&\frac{24i\nu \beta_0 -3 \nu^2 C_1 + \nu^4 C_2}{6}
1515: +\frac{3i\nu C_1 -6\beta_0-i\nu^3 C_2}{3r}
1516: -\frac{\nu^2 C_2}{r^2} +\frac{i\nu C_2}{r^3} +\frac{C_2}{2r^4},
1517: \label{e-NBl2}
1518: \end{eqnarray}
1519: with the (complex) constants $\beta_0$, $C_1$ and $C_2$ freely specifiable.
1520: In the case $\ell=3$
1521: \begin{eqnarray}
1522: \beta_2&=&\beta_0 \nonumber \\
1523: J_3(r)&=&\frac{60\beta_0 +3 i \nu C_1 + \nu^4 C_2}{180}+\frac{C_1}{10 r}
1524: -\frac{i \nu C_2}{6 r^3} -\frac{C_2}{4r^4}
1525: \nonumber \\
1526: U_3(r)&=&\frac{-60i\nu \beta_0 +3 \nu^2 C_1 - i \nu^5 C_2}{180}
1527: +\frac{2\beta_0}{r} +\frac{C_1}{2 r^2}
1528: -\frac{2\nu^2 C_2}{3 r^3} +\frac{5 i \nu C_2}{4 r^4}
1529: + \frac{C_2}{r^5}\nonumber \\
1530: W_{c3}(r)&=&\frac{60 i \nu \beta_0 -3 \nu^2 C_1 + i\nu^5 C_2}{15}
1531: +\frac{i\nu C_1 -2\beta_0+\nu^4 C_2}{3r}
1532: -\frac{i2\nu^3 C_2}{r^2} -\frac{4i\nu^2 C_2}{r^3}
1533: +\frac{5\nu C_2}{r^4}+\frac{3 C_2}{r^5}.
1534: \label{e-NBl3}
1535: \end{eqnarray}
1536:
1537: The news $N$ for the linearized wave is given by
1538: \begin{equation}
1539: N=\Re\left( e^{i\nu u}\lim_{r \rightarrow \infty}
1540: \left(-\frac{\ell(\ell+1)}{4}J_{\ell}-\frac{i\nu}{2} r^2 J_{\ell,r}\right)
1541: + e^{i\nu u}\beta_\ell \right) \sqrt{(\ell-1)\ell(\ell+1)(\ell+2)}
1542: \;{}_2R_{\ell m},
1543: \label{e-Nl}
1544: \end{equation}
1545: corresponding
1546: For the cases $\ell=$2 and 3, this gives
1547: \begin{equation}
1548: \ell=2:\;\;N=\Re\left(\frac{i\nu^3 C_2}{\sqrt{24}} e^{i\nu u}\right)
1549: \;{}_2R_{2m};\;\;\;
1550: \ell=3:\;\;N=\Re\left(\frac{-\nu^4 C_2}{\sqrt{30}} e^{i\nu u}\right)
1551: \;{}_2R_{3m}.
1552: \label{e-N}
1553: \end{equation}
1554: For the linearized case $\Psi=N_{,u}$, which gives
1555: \begin{equation}
1556: \ell=2:\;\;\Psi=\Re\left(\frac{-\nu^4 C_2}{\sqrt{24}} e^{i\nu u}\right)
1557: \;{}_2R_{2m};\;\;\;
1558: \ell=3:\;\;\Psi=\Re\left(\frac{-i\nu^5 C_2}{\sqrt{30}} e^{i\nu u}\right)
1559: \;{}_2R_{3m}.
1560: \end{equation}
1561:
1562: \subsection{Test specifications}
1563:
1564:
1565: Tests were run with the solution parameters $\nu=1$ and $m=0$
1566: for the cases $\ell=2$ and $\ell=3$, with
1567: \begin{eqnarray}
1568: C_1&=&3\cdot 10^{-6}\, , \quad C_2 =10^{-6}\, ,
1569: \quad \beta_0 =i\cdot 10^{-6} \quad (\ell=2) \\
1570: C_1&=&3\cdot 10^{-6}, \quad C_2 =i \cdot 10^{-6}\, ,
1571: \quad \beta_0 =i\cdot 10^{-6} \quad (\ell=3).
1572: \end{eqnarray}
1573: The inner worldtube boundary was placed at $r_0=2$ corresponding to
1574: a compactified radial coordinate $x_0=r_0/(R+r_0)\approx .1888$, where
1575: we have set the scale parameter $R=9$.
1576:
1577: For the convergence measurements, the $(\eta,\rho,x)$ grid consisted of $M^3$ points,
1578: with $M=100$ and $M=120$. The boundary of the circular patches were fixed at
1579: $\sqrt{\eta^2+\rho^2}=1.4$. The runs were stopped at $t=100$. The
1580: $L_\infty$ and $L_2$ error norms were computed on the North hemisphere, using
1581: the values from the North patch.
1582:
1583: Angular dissipation was applied only to the circular patch runs, with the
1584: dissipation coefficients $\epsilon_x=0.009$, $\epsilon_u=0.0009$,
1585: $\epsilon_Q=\epsilon_W=0.00001$. The weighting function ${\cal W}$ for
1586: application of the dissipation was taken to be a unit step function which
1587: vanishes for $\sqrt{\eta^2+\rho^2}\ge 1.3$.
1588:
1589: We present output data for the real parts of $J$, $N$
1590: and $N_\Psi$, For the $m=0$ case, these quantities correspond to
1591: a pure $\oplus$ polarization mode. For comparison purposes, we include
1592: results for the circular patch without dissipation and the original
1593: square patch treatment.
1594:
1595: \subsection{Test results for $J$}
1596:
1597: We first present test results for $J$, which is a typical metric quantity
1598: entering into the waveform calculation. Figure \ref{fig:InfErrJnorm} show the
1599: $L_\infty$ norm over the North hemisphere of the error ${\cal E}(J)$ vs the
1600: compactified radial coordinate $x$ at the end of the run at $t=100$ for the
1601: $\ell=2$ wave. The figure compares runs made with the circular patch method
1602: (dissipation applied) with runs without dissipation and runs with the original
1603: square patch method. The plots show that angular dissipation reduces the error.
1604: This will become more evident in the later test results for the news in which
1605: higher angular derivatives are involved. An important feature of the plots is
1606: that in all cases the error increases monotonically and takes it maximum value
1607: at ${\cal I}^+$ ($x=1$), as would be expected of the radial marching algorithm.
1608: This allows us to focus our error analysis on output at ${\cal I}^+$.
1609:
1610: Table \ref{tab:Conv_errorJ} gives the convergence rate of the error in $J$
1611: measured at ${\cal I}^+$ at various times during
1612: the $\ell=2$ run for the three methods
1613: shown in Fig.~\ref{fig:InfErrJnorm}. Clean second order convergence, measured
1614: either with an $L_2$ or $L_\infty$ norm, is indicated in all cases.
1615: The corresponding convergence rates for the $\ell=3$ runs are given in
1616: Table \ref{tab:Conv_errorJ_l3}.
1617: The $\ell =2$ runs are more discriminating because $|J|$ has a $\sin^2
1618: \theta$ dependence which peaks at the equator close to the interpatch
1619: interpolation, as opposed to the $\sin^2 \theta \cos \theta$ dependence of the
1620: $\ell=3$ case which vanishes at the equator. In the following we restrict our
1621: discussion to the $\ell=2$ case.
1622:
1623: The time dependence of the $L_2$ and $L_\infty$ errors in $J$ at ${\cal I}^+$
1624: for the circular patch run (with dissipation) is plotted in
1625: Fig.~\ref{fig:ErrJnorm}. The plots are based upon output at integer values of
1626: $t$, which samples the error at various phases during the underlying period
1627: $T=2\pi$. The errors for the two grids used in the convergence measurements are
1628: rescaled to the values for the finest $M=120$ grid, with the overlap again
1629: confirming clean convergence. The magnitude of the error is approximately
1630: $0.1\%$ the value of $J$. The $L_2$ error is
1631: smaller than the $L_\infty$ because the error is sharply peaked near the equator.
1632: This error
1633: pattern in the North hemisphere is exhibited in the snapshot of
1634: Fig.~\ref{fig:CrDissErrJSurf} at $t=100$. The profile is quite smooth - some of
1635: the apparent jaggedness near the edge is an artificial effect of the irregular
1636: pattern of grid points at the edge of the equator. The sharp spikes in the
1637: corresponding error snapshot for the circular run without dissipation shown in
1638: Figs.~\ref{fig:CrErrJSurfEG} illustrate the essential role of angular
1639: dissipation in guarding the Northern hemisphere from the interpolation error at
1640: the circular patch boundary. Such spikes are not apparent in the corresponding
1641: error snapshot for the square patch shown in Fig.~\ref{fig:SqErrJSurf}. The more
1642: regular square patch boundary does not require angular dissipation, although the
1643: resulting error is larger than for the circular patch with dissipation.
1644:
1645:
1646: \begin{figure}[htp] % figure placement: here, top, bottom, or page
1647: \centering
1648: \psfrag{xlabel}{x}
1649: \psfrag{ylabel}{${\cal E}(J)$}
1650: \includegraphics*[width=10cm]{Figure06}
1651: \caption{The $L_\infty$ error ${\cal E}(J)$ plotted vs $x$ at $t=100$ for
1652: runs with the circular patch method (with and without dissipation)
1653: and with the square patch method.}
1654: \label{fig:InfErrJnorm}
1655: \end{figure}
1656:
1657: \begin{figure}[htp] % figure placement: here, top, bottom, or page
1658: \centering
1659: \psfrag{xlabel}{t}
1660: \psfrag{ylabel}{${\cal E}(J)$}
1661: \includegraphics*[width=7cm]{Figure07}
1662: \includegraphics*[width=7cm]{Figure08}
1663: \caption{Plots of the error ${\cal E}(J)$ at ${\cal I}^+$ vs $t$,
1664: for the circular patch with dissipation measured with the
1665: $L_2$ norm (left plot) and the $L_\infty$ norm (right plot).
1666: The error for the $M=100$ grid is
1667: rescaled and overlaid on the error for the $M=120$ grid to exhibit
1668: the second order convergence. The smaller $L_2$ error
1669: indicates that the maximum error arises near the equator. }
1670: \label{fig:ErrJnorm}
1671: \end{figure}
1672:
1673:
1674: \begin{table}[htp]
1675: \begin{center}
1676: \begin{tabular}[c]{|c|c|c|c|c|c|c|c|c|}
1677: \hline
1678: Variable &
1679: circular patch &
1680: circular, no dissipation &
1681: square patch
1682:
1683: \\ \hline \hline
1684: %%%%%%
1685: ${\cal E}_{L_2}(J)_{t=1}$ & $2.01$ & $2.00$ & $2.01$
1686: \\ \hline
1687: %%%%%%
1688: %%%%%%
1689: ${\cal E}_{L_2}(J)_{t=10}$ & $2.01$ & $1.98$ & $2.00$
1690: \\ \hline
1691: %%%%%%
1692: %%%%%%
1693: ${\cal E}_{L_2}(J)_{t=90}$ & $2.00$ & $2.02$ & $2.02$
1694: \\ \hline
1695: %%%%%%
1696: %%%%%%
1697: ${\cal E}_{L_2}(J)_{t=100}$ & $1.92$ & $2.03$ & $2.00$
1698: \\ \hline
1699: %%%%%%
1700: %%%%%%
1701: ${\cal E}_{L_\infty}(J)_{t=1}$ & $2.01$ & $2.01$ & $2.01$
1702: \\ \hline
1703: %%%%%%
1704: %%%%%%
1705: ${\cal E}_{L_\infty}(J)_{t=10}$ & $1.95$ & $2.00$ & $1.99$
1706: \\ \hline
1707: %%%%%%
1708: %%%%%%
1709: ${\cal E}_{L_\infty}(J)_{t=90}$ & $2.07$ & $1.96$ & $2.00$
1710: \\ \hline
1711: %%%%%%
1712: %%%%%%
1713: ${\cal E}_{L_\infty}(J)_{t=100}$ & $1.92$ & $2.01$ & $1.99$
1714: \\ \hline
1715: %%%%%%
1716: \end{tabular}
1717: \caption{Convergence rates of the error ${\cal E}(J)$ at ${\cal I}^+$
1718: for the $\ell=2$ run,
1719: measured at times $t=1$, $t=10$, $t=90$, and $t=100$.
1720: }
1721: \label{tab:Conv_errorJ}
1722: \end{center}
1723: \end{table}
1724:
1725: \begin{table}[htp]
1726: \begin{center}
1727: \begin{tabular}[c]{|c|c|c|c|c|c|c|c|c|}
1728: \hline
1729: Variable &
1730: circular patch &
1731: circular, no dissipation &
1732: square patch
1733:
1734: \\ \hline \hline
1735: %%%%%%
1736: ${\cal E}_{L_2}(J)_{t=1}$ & $2.02$ & $2.01$ & $2.01$
1737: \\ \hline
1738: %%%%%%
1739: %%%%%%
1740: ${\cal E}_{L_2}(J)_{t=10}$ & $2.00$ & $2.00$ & $2.00$
1741: \\ \hline
1742: %%%%%%
1743: %%%%%%
1744: ${\cal E}_{L_2}(J)_{t=90}$ & $2.03$ & $2.02$ & $2.02$
1745: \\ \hline
1746: %%%%%%
1747: %%%%%%
1748: ${\cal E}_{L_2}(J)_{t=100}$ & $2.05$ & $2.00$ & $2.01$
1749: \\ \hline
1750: %%%%%%
1751: %%%%%%
1752: ${\cal E}_{L_\infty}(J)_{t=1}$ & $2.02$ & $2.02$ & $2.02$
1753: \\ \hline
1754: %%%%%%
1755: %%%%%%
1756: ${\cal E}_{L_\infty}(J)_{t=10}$ & $1.99$ & $1.99$ & $2.00$
1757: \\ \hline
1758: %%%%%%
1759: %%%%%%
1760: ${\cal E}_{L_\infty}(J)_{t=90}$ & $2.02$ & $2.02$ & $2.04$
1761: \\ \hline
1762: %%%%%%
1763: %%%%%%
1764: ${\cal E}_{L_\infty}(J)_{t=100}$ & $2.00$ & $2.00$ & $1.99$
1765: \\ \hline
1766: %%%%%%
1767: \end{tabular}
1768: \caption{Convergence rates of the error ${\cal E}(J)$ at ${\cal I}^+$
1769: for the $\ell =3$ run,
1770: measured at times $t=1$, $t=10$, $t=90$, and $t=100$.
1771: }
1772: \label{tab:Conv_errorJ_l3}
1773: \end{center}
1774: \end{table}
1775:
1776:
1777: \begin{figure}[htp] % figure placement: here, top, bottom, or page
1778: \centering
1779: \includegraphics*[width=14cm]{Figure09}
1780: \caption{Surface plot of the error in $J$ at $t=100$ for the
1781: circular patch run (with dissipation).}
1782: \label{fig:CrDissErrJSurf}
1783: \end{figure}
1784:
1785: \begin{figure}[htp] % figure placement: here, top, bottom, or page
1786: \centering
1787: \includegraphics*[width=14cm]{Figure10}
1788: \caption{Surface plot of the error in $J$ at $t=100$ for the circular
1789: patch run without dissipation.}
1790: \label{fig:CrErrJSurfEG}
1791: \end{figure}
1792:
1793:
1794: \begin{figure}[htp] % figure placement: here, top, bottom, or page
1795: \centering
1796: \includegraphics*[width=14cm]{Figure11}
1797: \caption{Surface plot of the error in $J$ at $t=100$ for the square
1798: patch run.}
1799: \label{fig:SqErrJSurf}
1800: \end{figure}
1801:
1802: \clearpage
1803:
1804:
1805: \subsection{Test results for the news}
1806:
1807: We now compare test results for the news function in terms of a direct
1808: calculation of $N$ via (\ref{eq:news}) and a calculation of $N_\Psi$ via
1809: (\ref{eq:npsi}) using the Weyl component $\Psi$ given in (\ref{eq:psia}). We
1810: restrict the discussion to the $\ell=2$ runs which are more challenging than
1811: $\ell=3$ with respect to problems near the equator. Tables \ref{tab:Conv_errorN}
1812: and \ref{tab:Conv_errorNewsPsi} give the convergence rates of the $L_2$ and
1813: $L_\infty$ errors in $N$ and $N_\Psi$ measured at various times for runs with
1814: the circular patch (with dissipation), the circular patch without dissipation
1815: and the original square patch methods. At the final run time $t=100$,
1816: measurements for all cases show clean second order convergence, although there
1817: is a small departure in the $N_\Psi$ rates at early times.
1818:
1819: The plots of the $L_2$ error vs time for the circular patch runs in
1820: Fig.~\ref{fig:ErrNewsnorm} show little difference in the time behavior between
1821: $N$ and $N_\Psi$, although the error in $N_\Psi$ is slightly smaller. The
1822: $L_\infty$ errors measured at the end of the runs on the $M=120$ grid are given
1823: in Table \ref{tab:InfNull_norminf} for the circular patch, the circular patch
1824: without dissipation and the square patch. The best results are obtained for the
1825: circular patch, which shows an $\approx 30\%$ improvement over the original
1826: square patch treatment. The results also show the essential improvement due to
1827: the use of angular dissipation. For the circular patch, the error in $N_\Psi$
1828: was $\approx 24\%$ smaller than the error in $N$ at the end of the run. But it
1829: is also clear from the plots of the $L_2$ error Fig.~\ref{fig:ErrNewsnorm} that
1830: this ratio depends when and where this ratio is taken. At the equator where the
1831: news takes its maximum value, its modulus for this test is $|N_{analytic}|
1832: \approx 8\times 10^{-8}$. At the end of run, the corresponding fractional errors
1833: in $N_\Psi$ and $N$ are $\approx 4\%$ for averaged values and $\approx 9\%$ for
1834: the maximum errors at the equator.
1835:
1836: Surface plots of the error in $N$ and $N_\Psi$ at the end of the run are given
1837: in Figs.~\ref{fig:CrDissErrNewsSurf} - \ref{fig:SqnoDErrNPsi4Surf} for the
1838: circular and square patches. The lack of sharp spikes in the errors for the
1839: circular patches shows the effectiveness of applying angular dissipation. There
1840: is slightly more jaggedness near the equator for the circular vs square patch
1841: errors, but this is overbalanced by the relative smallness of the circular patch
1842: error.
1843:
1844: \begin{table}[htp]
1845: \begin{center}
1846: \begin{tabular}[c]{|c|c|c|c|c|c|c|c|c|}
1847: \hline
1848: Variable &
1849: circular patch &
1850: circular, no dissipation&
1851: square patch
1852:
1853: \\ \hline \hline
1854: %%%%%%
1855: ${\cal E}_{L_2}(N)_{t=1}$ & $2.05$ & $2.05$ & $2.05$
1856: \\ \hline
1857: %%%%%%
1858: %%%%%%
1859: ${\cal E}_{L_2}(N)_{t=10}$ & $2.05$ & $2.05$ & $2.04$
1860: \\ \hline
1861: %%%%%%
1862: %%%%%%
1863: ${\cal E}_{L_2}(N)_{t=90}$ & $2.04$ & $2.04$ & $2.01$
1864: \\ \hline
1865: %%%%%%
1866: %%%%%%
1867: ${\cal E}_{L_2}(N)_{t=100}$ & $2.01$ & $2.07$ & $2.05$
1868: \\ \hline
1869: %%%%%%
1870: %%%%%%
1871: ${\cal E}_{L_\infty}(N)_{t=1}$ & $2.04$ & $2.04$ & $2.04$
1872: \\ \hline
1873: %%%%%%
1874: %%%%%%
1875: ${\cal E}_{L_\infty}(N)_{t=10}$ & $2.04$ & $1.99$ & $2.04$
1876: \\ \hline
1877: %%%%%%
1878: %%%%%%
1879: ${\cal E}_{L_\infty}(N)_{t=90}$ & $2.01$ & $2.01$ & $2.06$
1880: \\ \hline
1881: %%%%%%
1882: %%%%%%
1883: ${\cal E}_{L_\infty}(N)_{t=100}$ & $1.98$ & $2.00$ & $1.93$
1884: \\ \hline
1885: %%%%%%
1886: \end{tabular}
1887: \caption{Convergence rates of the error ${\cal E}(N)$,
1888: measured at $t=1$, $t=10$, $t=90$, and $t=100$.
1889: }
1890: \label{tab:Conv_errorN}
1891: \end{center}
1892: \end{table}
1893:
1894: \begin{table}[htp]
1895: \begin{center}
1896: \begin{tabular}[c]{|c|c|c|c|c|c|c|c|c|}
1897: \hline
1898: Variable &
1899: circular patch &
1900: circular, no dissipation&
1901: square patch
1902:
1903: \\ \hline \hline
1904: %%%%%%
1905: ${\cal E}_{L_2}(N_{\Psi})_{t=1}$ & $2.11$ & $2.10$ & $2.11$
1906: \\ \hline
1907: %%%%%%
1908: %%%%%%
1909: ${\cal E}_{L_2}(N_{\Psi})_{t=10}$ & $2.13$ & $2.13$ & $2.11$
1910: \\ \hline
1911: %%%%%%
1912: %%%%%%
1913: ${\cal E}_{L_2}(N_{\Psi})_{t=90}$ & $2.09$ & $2.09$ & $2.08$
1914: \\ \hline
1915: %%%%%%
1916: %%%%%%
1917: ${\cal E}_{L_2}(N_{\Psi})_{t=100}$ & $2.02$ & $1.98$ & $2.00$
1918: \\ \hline
1919: %%%%%%
1920: %%%%%%
1921: ${\cal E}_{L_\infty}(N_{\Psi})_{t=1}$ & $2.08$ & $2.08$ & $2.08$
1922: \\ \hline
1923: %%%%%%
1924: %%%%%%
1925: ${\cal E}_{L_\infty}(N_{\Psi})_{t=10}$ & $2.09$ & $2.05$ & $2.10$
1926: \\ \hline
1927: %%%%%%
1928: %%%%%%
1929: ${\cal E}_{L_\infty}(N_{\Psi})_{t=90}$ & $2.05$ & $2.00$ & $2.06$
1930: \\ \hline
1931: %%%%%%
1932: %%%%%%
1933: ${\cal E}_{L_\infty}(N_{\Psi})_{t=100}$ & $1.98$ & $2.01$ & $1.93$
1934: \\ \hline
1935: %%%%%%
1936: \end{tabular}
1937: \caption{Convergence rates of the error ${\cal E}(N_{\Psi})$,
1938: measured at $t=1$, $t=10$, $t=90$, and $t=100$.
1939: }
1940: \label{tab:Conv_errorNewsPsi}
1941: \end{center}
1942: \end{table}
1943:
1944: \begin{table}[htp]
1945: \begin{center}
1946: \begin{tabular}[c]{|c|c|c|c|}
1947: \hline
1948: Variable &
1949: circular patch &
1950: circular, no dissipation&
1951: square patch
1952:
1953: \\ \hline \hline
1954: %%%%%
1955: ${\cal E}_{L_\infty}(N)$ & $2.247\times 10^{-9}$ & $3.325\times 10^{-9}$ & $2.897\times 10^{-9}$
1956: \\ \hline
1957: %%%%%
1958: %%%%%
1959: ${\cal E}_{L_\infty}(N_{\Psi})$ & $1.706\times 10^{-9}$ & $2.747\times 10^{-9}$ & $2.315\times 10^{-9}$
1960: \\ \hline
1961: %%%%%
1962: \end{tabular}
1963: \caption{The values of the $L_\infty$ errors in $N$ and $N_\Psi$
1964: measured at $t=100$,}
1965: \label{tab:InfNull_norminf}
1966: \end{center}
1967: \end{table}
1968:
1969: \begin{figure}[htp] % figure placement: here, top, bottom, or page
1970: \centering
1971: \begin{psfrags}
1972: \psfrag{xlabel}{t}
1973: \psfrag{ylabel}{${\cal E}(N)$}
1974: \includegraphics*[width=7cm]{Figure12}
1975: \end{psfrags}
1976: \begin{psfrags}
1977: \psfrag{xlabel}{t}
1978: \psfrag{ylabel}{${\cal E}(N_{\Psi})$}
1979: \includegraphics*[width=7cm]{Figure13}
1980: \end{psfrags}
1981: \caption{Plots of the $L_2$ errors vs t for
1982: $N$ (left plot) and $N_\Psi$ (right plot) for the circular patch
1983: runs. The plots for the $M=100$ grid are rescaled to
1984: the $M=120$ grid. The plots are based upon output at at integer
1985: values of $t$.
1986: }
1987: \label{fig:ErrNewsnorm}
1988: \end{figure}
1989:
1990: \begin{figure}[htp] % figure placement: here, top, bottom, or page
1991: \centering
1992: \includegraphics*[width=14cm]{Figure14}
1993: \caption{Surface plot of the error in $N$ at $t=100$ for the
1994: circular patch.}
1995: \label{fig:CrDissErrNewsSurf}
1996: \end{figure}
1997:
1998:
1999: \begin{figure}[htp] % figure placement: here, top, bottom, or page
2000: \centering
2001: \includegraphics*[width=14cm]{Figure15}
2002: \caption{Surface plot of the error in $N$ at $t=100$ for the
2003: square patch. }
2004: \label{fig:SqErrNewsSurf}
2005: \end{figure}
2006:
2007: \begin{figure}[htp] % figure placement: here, top, bottom, or page
2008: \centering
2009: \includegraphics*[width=14cm]{Figure16}
2010: \caption{Surface plot of the error in $N_{\Psi}$ at $t=100$
2011: for the circular patch.}
2012: \label{fig:CrDissErrNPsi4Surf}
2013: \end{figure}
2014:
2015:
2016: \begin{figure}[htp] % figure placement: here, top, bottom, or page
2017: \centering
2018: \includegraphics*[width=14cm]{Figure17}
2019: \caption{Surface plot of the error in $N_{\Psi}$ at $t=100$
2020: for the square patch.}
2021: \label{fig:SqnoDErrNPsi4Surf}
2022: \end{figure}
2023:
2024: \clearpage
2025:
2026:
2027: \section{Conclusion}
2028: \label{sec:concl}
2029:
2030: We have proposed two new methods for enhancing the accuracy of CCE, One is a
2031: numerical method modifying the stereographic patches used in the
2032: characteristic evolution code to conform to the circular patch boundaries as
2033: used in meteorology~\cite{browning}. The other is a geometrical method that
2034: bases the waveform on the limiting behavior at ${\cal I}^+$ of the Weyl tensor
2035: component $\Psi_4$ rather than the news function.
2036:
2037: We have used a scalar wave testbed to compare the circular patch method against
2038: the cubed sphere method, which is also extensively used in
2039: meteorology~\cite{ronchi}. We found, for equivalent computational expense, that
2040: the cubed-sphere method has an edge in accuracy over the stereographic method.
2041: The cubed-sphere error in the scalar field ${\cal E}(\Phi)$ is $\approx
2042: \frac{1}{3}$ the stereographic error but that the advantage is smaller for the
2043: higher derivatives required in gravitational waveform extraction. The
2044: cubed-sphere error ${\cal E}(\bar \eth \eth^2\Phi)$ is only $\approx
2045: \frac{4}{5}$ the stereographic error. An advantage of the stereographic approach
2046: is its relative programming simplicity. But as originally pointed out
2047: in~\cite{ronchi}, and demonstrated recently for the case of a characteristic
2048: evolution code~\cite{roberto}, once all the necessary infrastructure for
2049: interpatch communication is in place, the shared boundaries of the cubed-sphere
2050: approach admit a highly scalable algorithm for parallel architectures.
2051:
2052: We used the circular patch stereographic code to compare waveform extraction
2053: in a linearized wave test directly via the Bondi news function $N$
2054: and its counterpart $N_\Psi$
2055: constructed from the Weyl curvature. For this purpose, we were able to
2056: successfully implement a new form of angular dissipation in the characteristic
2057: evolution code, which otherwise would be prone to high frequency error
2058: introduced by the irregular way a circular boundary cuts through a square grid.
2059: Our test results show that this dissipation works: the
2060: resulting error in the waveforms and metric quantities is smooth. In addition,
2061: the extensive analytic and numerical manipulations carried out to compute the
2062: limiting behavior of the Weyl curvature was demonstrated to yield second order
2063: accurate results for $N_\Psi$.
2064:
2065: In the linearized tests presented here, neither $N$ nor $N_\Psi$ was a clear
2066: winner. We already knew that the original news module based upon a square
2067: stereographic patch worked well in the linear regime. The news $N$ calculated on
2068: a circular patch had lower error than that on a square patch but only by a
2069: $\approx 30\%$ factor. In turn, the news calculated via $N_\Psi$ on the circular
2070: patch had a lower error than $N$ on the circular patch by a $\approx 24\%$
2071: factor. Weyl tensor extraction is slightly more accurate than news function
2072: extraction, even though there are many more terms involved.
2073:
2074: All errors were second order convergent. However, while there was a small
2075: fractional error $\approx .1\%$ in metric quantities such as $J$, the
2076: corresponding averaged error in the $N_\Psi$ and $N$ was $\approx 4\%$
2077: for the circular patch runs and the maximum error at the equator was
2078: $\approx 9\%$. These errors did not vary appreciably ($\approx 30\%$)
2079: with the choice of discretization method, i.e. circular patch or
2080: square patch. They reflect the intrinsic difficulty in extracting waveforms due to
2081: the delicate cancellation of leading order terms in the underlying metric and
2082: connection when computing $O(1/r)$ quantities such as $\Psi_4$. The excellent
2083: accuracy that we find for the metric suggests that perturbative waveform
2084: extraction must suffer the same difficulty. In that case it is just less obvious
2085: how to quantify the errors. The delicate issues involved at ${\cal I}^+$ have
2086: been shown to have counterparts in extraction on a finite
2087: worldtube~\cite{lehnmor}.
2088:
2089: Waveforms are not easy to extract accurately. However, the convergence of our
2090: error measurements is a positive sign that higher order finite difference
2091: approximations might supply the accuracy that is needed for realistic
2092: astrophysical applications. Whether the advantages the new methods proposed here
2093: prove to be significant will depend upon the results of future application in
2094: the nonlinear regime.
2095:
2096: \centerline{\bf Acknowledgments}
2097:
2098: We thank Thomas Maedler for checking the calculations in Sec.~\ref{sec:weyl} and
2099: G.~L. Browning for correspondence concerning the application of stereographic
2100: patches in computational fluid dynamics. NTB thanks Max-Planck-Institut f\" ur
2101: Gravitationsphysik, Albert-Einstein-Institut for hospitality; BS thanks
2102: University of South Africa for hospitality; and MCB thanks University of
2103: Pittsburgh for hospitality. We have benefited from the use of the Cactus
2104: Computational Toolkit (http://www.cactuscode.org). Computer time was provided by
2105: the Pittsburgh Supercomputing Center through a TeraGrid Wide Roaming Access
2106: Computational Resources Award, and we owe special thanks to R. G\'{o}mez for his
2107: assistance. This work was supported by the Sherman Fairchild Foundation and the
2108: National Science Foundation under grants PHY-061459 and PHY-0652995 to the
2109: California Institute of Technology; the National Science Foundation grant
2110: PH-0553597 to the University of Pittsburgh; and by the National Research
2111: Foundation, South Africa, under GUN 2075290.
2112:
2113: \begin{thebibliography}{40}
2114:
2115: \bibitem{bondi}
2116: H. Bondi, M.J.G. van der Burg and A.W.K. Metzner,
2117: {\em Proc. R. Soc. A} {\bf 269} 21, 1962.
2118:
2119: \bibitem{sachs}
2120: R.K. Sachs, {\em Proc. R. Soc. A} {\bf 270} 103, 1962.
2121:
2122: \bibitem{Penrose}
2123: R. Penrose, {\em Phys. Rev. Letters}, {\bf10} 66, 1963.
2124:
2125: \bibitem{isaac}
2126: R.A. Isaacson, J.S. Welling and J. Winicour, {\em J. Math. Phys.} {\bf
2127: 24} 1824, 1983.
2128:
2129: \bibitem{highp} N.~T. Bishop, R.~G\'{o}mez, L. Lehner, M. Maharaj and
2130: J. Winicour, {\it Phys. Rev. D},{\bf 56} 6298 (1997).
2131:
2132: \bibitem{cce}
2133: N.T. Bishop, R. G\'{o}mez, L. Lehner, and J. Winicour, {\em Phys. Rev. D},
2134: {\bf 54} 6153, 1996.
2135:
2136: \bibitem{hypb} H. Friedrich, ``Conformal Einstein evolution''
2137: {\it Lecture Notes in Physics} {\bf 604}, 1 (2002).
2138:
2139: \bibitem{Hhprbloid} S. Husa, ``Numerical Relativity with the conformal
2140: field equations'', {\it Lecture Notes in Physics} {\bf 617}, 159 (2003).
2141:
2142: \bibitem{joerg} J. Frauendiener, ``Conformal infinity'',
2143: {\it Living Rev. Relativity} {\bf 7}, 253 (2004).
2144:
2145: \bibitem{ccm} N.T. Bishop, R. G\'{o}mez, L. Lehner, B. Szil\'{a}gyi,
2146: J. Winicour and R. A. Isaacson,
2147: ``Cauchy-Characteristic Matching'', in B Iyer and B Bhawal (Eds.),
2148: {\it Black Holes, Gravitational Radiation and the Universe},
2149: Kluwer Academic Publishers, Dordrecht, 1998
2150:
2151: \bibitem{livccm} Winicour, J. (2005), ``Characteristic evolution and matching'',
2152: {\it Living Rev. Relativity} {\bf 8}, 10.
2153:
2154: \bibitem{harm} B. Szil{\'{a}}gyi and J. Winicour,
2155: ``Well-posed initial-boundary evolution in general relativity'',
2156: {\it Phys. Rev. D} {\bf 68} 041501 (2003).
2157:
2158: \bibitem{gk} B. Gustaffson and H.-O. Kreiss,
2159: ``Boundary conditions for time dependent problems with
2160: an artificial boundary'', {\it J. Comp. Phys.}
2161: {\bf 30}, 333--351 (1979).
2162:
2163: \bibitem{friednag} H. Friedrich, H. and G. Nagy,
2164: {\it Commun. Math. Phys.}, {\bf 201}, 619 (1999).
2165:
2166: \bibitem{NP}
2167: E.T. Newman and R. Penrose, {\em J. Math. Phys.} {\bf 3}, 566, 1962.
2168:
2169: \bibitem{psiocit} O. Rinne, L. Lindblom and M.A. Scheel, ``Testing outer
2170: boundary treatments for the Einstein equations'' {\it Class. Quantum Grav.}
2171: {\bf 24}, 4053 (2007).
2172:
2173: \bibitem{regwh} T. Regge and J.~A. Wheeler, {\it Phys. Rev.}
2174: {\bf 108}, 1063 (1957).
2175:
2176: \bibitem{zeril} F. Zerilli, {\it Phys. Rev. Lett.}
2177: {\bf 24}, 737 (1970).
2178:
2179: \bibitem{nagrez} A. Nagar and L. Rezzolla, {\it Class. Quantum Grav.}
2180: {\bf 22}, R167 (2005).
2181:
2182: \bibitem{bakcamluo} J. Baker, M. Campanelli, C.O. Lousto and R. Takahashi
2183: ``Modeling gravitational radiation from coalescing binary black holes'',
2184: {\it Phys.Rev. D}, {\bf 65}, 124012--124034 (2002).
2185:
2186: \bibitem{pretlet} F. Pretorius, {\it Phys. Rev. Lett.} {\bf 95}, 121101 (2005).
2187:
2188: \bibitem{camluomarzlo} M. Campanelli, C. O. Lousto, P. Marronetti
2189: and Y. Zlochower, ``Accurate Evolutions of Orbiting Black-Hole Binaries
2190: Without Excision'', {\it Phys.Rev.Lett.} {\bf 96} 111101 (2006)
2191:
2192: \bibitem{bakcenchokopvme} J.~G. Baker, J. Centrella, D-I. Choi,
2193: M. Koppitz and J. van Meter,
2194: ``Binary black hole merger dynamics and waveforms'',
2195: {\it Phys.Rev. D} {\bf 73}, 104002 (2006)
2196:
2197: \bibitem{lehnmor} L. Lehner and O.~M. Moreschi,
2198: ``Dealing with delicate issues in waveform calculations'', gr-qc:0706.1319.
2199:
2200: \bibitem{babiuc05}
2201: M. Babiuc, B. Szil\'{a}gyi, I. Hawke, and Y. Zlochower,
2202: ``Gravitational wave extraction based on Cauchy-characteristic extraction
2203: and characteristic evolution'',
2204: {\it Class. Quantum Grav.} {\bf 22} 5089 (2005)
2205:
2206: \bibitem{partbh} N.~T. Bishop, R. G\'{o}mez, S. Husa, L. Lehner, J. Winicour
2207: ``A numerical relativistic model of a massive particle in orbit near a Schwarzschild black hole'',
2208: {\it Phys.Rev. D} {\bf 68}, 084015 (2003)
2209:
2210: \bibitem{pretlehn} F. Pretorius and L. Lehner,
2211: {\it J. Comput. Phys.} {\bf 198}, 10 (2004)
2212:
2213: \bibitem{browning} G. L. Browning, J. J. Hack and P. N. Swarztrauber
2214: ``A Comparison of Three Numerical Methods for Solving Differential Equations on the Sphere'',
2215: {\it Monthly Weather Review} {\bf 117}, 10582 (1989).
2216:
2217: \bibitem{ronchi} C. Ronchi, R. Iacono and P.~S. Paolucci,
2218: ``The ``cubed sphere'': a new method for the solution of partial differential
2219: equations on the sphere'', {\it J. Comput. Phys.} {\bf 124}, 93--114 (1996)
2220:
2221: \bibitem{eth}
2222: R. G\'{o}mez, L. Lehner, P. Papadopoulos and J. Winicour,
2223: ``The eth formalism in numerical relativity'',
2224: {\it Class. Quantum Grav.} {\bf 14} 977, 1997.
2225:
2226: \bibitem{Thornburgah} J. Thornburg
2227: ``A fast apparent horizon finder for three-dimensional Cartesian grids
2228: in numerical relativity'',
2229: {\it Class. Quantum Grav.} {\bf 21} 743-766 (2004)
2230:
2231: \bibitem{reisswig} C. Reisswig, N.T. Bishop, C.W. Lai, J. Thornburg
2232: and B. Szil\'{a}gyi
2233: ``Characteristic Evolutions in Numerical relativity using Six Angular Patches'',
2234: {\it Class. Quantum Grav.} {\bf 24} S237-S339 (2007)
2235:
2236: \bibitem{roberto}
2237: R. G\'{o}mez, W. Barreto, and S. Frittelli,
2238: ``A framework for large-scale relativistic simulations in the
2239: characteristic approach'',
2240: {\it Phys. Rev. D} {\bf 76} 124029--124050 (2007).
2241:
2242: \bibitem{kreisoligd} H-O. Kreiss and J. Oliger,
2243: ``Methods for the approximate solution of time dependent problems'',
2244: Global Atmospheric Research Program, {\bf Publication No. 10},
2245: World Meteorological Organization,
2246: Case Postale No. 1, CH-1211 Geneva 20, Switzerland.
2247:
2248: \bibitem{luisdis} L. Lehner,
2249: ``A dissipative algorithm for wave-like equations in the characteristic formulation'',
2250: {\it J. Comput. Phys.} {\bf 149}, 59--74 (1999).
2251:
2252: \bibitem{newp}
2253: E.T. Newman and R. Penrose, J. Math. Phys. {\bf 7}, 863 (1966).
2254:
2255: \bibitem{newt}
2256: J. Winicour, {\em J. Math. Phys.} {\bf 24} 1193, 1983.
2257:
2258: \bibitem{nullinf}
2259: J. Winicour, {\em J. Math. Phys.} {\bf 25} 2506, 1984.
2260:
2261: \bibitem{gomezfo} R. G{\'{o}}mez,
2262: ``Gravitational waveforms with controlled accuracy'',
2263: {\it Phys. Rev. D} {\bf 64}, 1--8 (2001).
2264:
2265: \bibitem{tam}
2266: L.A. Tamburino and J. Winicour, {\em Phys. Rev.} {\bf 150} 1039, 1966.
2267:
2268: \bibitem{quad}
2269: J. Winicour, {\em Gen. Rel. and Grav.} {\bf 19} 281 (1987).
2270:
2271: \bibitem{BS-lin}
2272: N.T. Bishop, Class. Quantum Grav. {\bf 22} 2393 (2005).
2273:
2274: \bibitem{golm}
2275: J.N. Goldberg, A.J. MacFarlane, E.T. Newman, F. Rohrlich and
2276: E.C.G. Sudarshan, J. Math. Phys. {\bf 8}, 2155 (1967).
2277:
2278: \bibitem{mod}
2279: Y.~Zlochower, R. G\'{o}mez, S.~Husa, L.~Lehner and J.~Winicour,
2280: ``Mode coupling in the nonlinear response of black holes,''
2281: {\it Phys. Rev. D} {\bf 68}, 084014 (2003).
2282:
2283: \end{thebibliography}
2284:
2285: \end{document}
2286: