gr-qc0307067/z48.tex
1: %\documentclass[a4paper]{article}
2: %\def \Box {\Delta}
3: \def \tr {{\rm tr}}
4: %\begin{document}
5: %\documentstyle[preprint,aps,prd]{revtex}
6: %\documentstyle[10pt]{revtex}
7: %\tighten
8: \documentclass[twocolumn,showpacs,preprintnumbers,amsmath,amssymb]{revtex4}
9: %\documentclass[showpacs,preprintnumbers,amsmath,amssymb,10pt]{revtex4}
10: %\draft
11: 
12: %\usepackage{amstex}   % useful for coding complex math
13: %\mathindent\parindent % needed in case "Amstex" is used
14: \usepackage{epsf}
15: 
16: \begin{document}
17: \title{A symmetry-breaking mechanism for the Z4 general-covariant evolution
18: system}
19: \author{C.~Bona, T.~Ledvinka, C.~Palenzuela and M.~\v Z\' a\v cek
20: %}\affiliation{
21: \\ Departament de Fisica, Universitat de les Illes Balears,
22: \\     Ctra de Valldemossa km 7.5, 07071 Palma de Mallorca, Spain}
23: %\address{Institute of Theoretical Physics, Faculty of Mathematics
24: %and Physics, Charles University, V Hole\v{s}ovi\v{c}k\'ach 2, 180
25: %00 Prague 8, Czech Republic}
26: 
27: \begin{abstract}
28: The general-covariant Z4 formalism is further analyzed. The gauge conditions are
29: generalized with a view to Numerical Relativity applications and the conditions
30: for obtaining strongly hyperbolic evolution systems are given both at the first
31: and the second order levels. A symmetry-breaking mechanism is proposed that
32: allows one, when applied in a partial way, to recover previously proposed
33: strongly hyperbolic formalisms, like the BSSN and the Bona-Mass\'o ones. When
34: applied in its full form, the symmetry breaking mechanism allows one to recover
35: the full five-parameter family of first order KST systems. Numerical codes based in the
36: proposed formalisms are tested. A robust stability test is provided by evolving
37: random noise data around Minkowski space-time. A strong field test is provided by
38: the collapse of a periodic background of plane gravitational waves, as described
39: by the Gowdy metric.
40: \end{abstract}
41: 
42: 
43: \maketitle
44: %\baselineskip=32pt
45: \section{Introduction}
46: 
47: 
48: 
49: The waveform emitted in the inspiral and merger of a relativistic binary is a
50: theoretical input crucial to the success of the laser interferometry gravitational
51: laboratories \cite{Ligo,Virgo,Geo,Tama}. Although the regular orbiting phase can be
52: treated with good accuracy by well-known analytical perturbation methods, the
53: later phases belong clearly to the strong field regime of either a black hole
54: or a neutron star collision, so that a computational approach is mandatory.
55: This kind of computational effort has been the objective of the Binary Black
56: Hole (BBH) Grand Challenge \cite{GC} and other world wide collaborations. The
57: resulting numerical codes are based on the so called ADM formalism \cite{ADM62}
58: for the Einstein field equations, where only a subset of the equations are
59: actually used for evolution whereas the remaining ones are considered as constraints to
60: be imposed on the initial data only (free evolution approach \cite{Cent80}).
61: 
62: It is clear that, by taking the constraints out of the evolution system, one is extending
63: the solution space. This extension is the crucial step that opened the way to
64: new hyperbolic formalisms after the seminal work of Y. Choquet-Bruhat and
65: T. Ruggeri \cite{CR83}, using the constraints to modify the evolution
66: system in many different ways \cite{BM92,BM95,FR94,FR96,SN95,BS99,AY99,Hern00,KST01,ST02},
67: even taking additional derivatives \cite{CR83,AACY95,Frie96}. These formalisms can be
68: interpreted as providing many non-equivalent ways of extending the solution space
69: of Einstein's equations with at least one common feature: constraint equations
70: are left out of the final evolution system.
71: 
72: This means that the resulting systems do have an extended solution space, which includes
73: constraint-violating solutions in addition to the ones verifying the original Einstein's
74: equations. As far as the constraints are first integrals of the extended evolution system,
75: Einstein's solutions could be computed by solving the constraints equations only for the initial
76: data (free evolution). But, unless some enforcing mechanism is used during the subsequent
77: time evolution, numerical errors will activate constraint-violating modes. Numerical
78: simulations can deal with such modes, at least when the deviations from Einstein's solutions
79: are moderate. But it happens that large deviations are usually associated with instabilities
80: of constraint-violating modes, leading to code crashing.
81: 
82: In particular, numerical codes based on these new hyperbolic formalisms happen to be quite
83: intolerant to violations of the Hamiltonian constraint in the initial data.
84: This is a serious drawback if one is planning to use the results of the
85: analytical approximation to the regular orbiting phase of a binary system as
86: initial data for a numerical simulation of the final ringdown and merger. The
87: numerical code will crash before any template for the gravitational wave emission
88: could be extracted. This is because the analytical data are just a good
89: approximation, so that the energy constraint does not hold exactly and the code
90: is intolerant to that kind of ``off-shell" initial data. Although there can be other
91: options, we claim that a numerical code tolerant to constraint violation would be
92: undoubtedly the best alternative.
93: 
94: There have been some attempts in that direction. In Ref.
95: \cite{BFHR99} the technique of Lagrange multipliers was used for
96: including the constraints into the dynamical system. The extended
97: system includes the Lagrange multipliers as additional dynamical
98: fields ($\lambda$-system). A further step along that direction is
99: given in Ref. \cite{SY}, where the extended system is "adjusted"
100: by further combining the evolution equations with the constraints,
101: including then many extra arbitrary parameters. In both cases, the
102: goal is to enforce the constraints in a dynamical way. One can
103: monitor the errors by looking at the "subsidiary system", which
104: can be derived from Bianchi identities and can be interpreted as
105: the evolution system for constraint deviations. Parameters are
106: adjusted in a way such that the characteristic speeds of the
107: subsidiary system are either real and non-zero (so that the
108: corresponding errors propagate away to the boundaries) or they
109: have the right sign in the imaginary part to enforce damping
110: (instead of exploding) the constraint deviations.
111: 
112: A simpler option (but not the only one, see for instance
113: \cite{GGKM03}) for including the constraints into the evolution system 
114: would be the general-covariant extension of the Einstein field equations proposed
115: recently (Z4 system) \cite{Z4}:
116: \begin{equation}\label{EinsteinZ4}
117:   R_{\mu \nu} + \nabla_{\mu} Z_{\nu} + \nabla_{\nu} Z_{\mu} =
118:   8\; \pi\; (T_{\mu \nu} - \frac{1}{2}\, T\; g_{\mu \nu})~,
119: \end{equation}
120: so that the full set of dynamical fields {consists of} the pair
121: $\{g_{\mu\nu},Z_\mu\}$. The solutions of the original field
122: equations can then be recovered by imposing the algebraic constraint
123: \begin{equation}\label{Zis0}
124:   Z_{\mu} = 0
125: \end{equation}
126: and the evolution of this constraint is subject to the
127: linear homogeneous equation
128: \begin{equation}\label{WaveZis0}
129:   \Box\; Z_{\mu} + R_{\mu\nu} Z^\nu = 0~,
130: \end{equation}
131: which can be easily obtained from (\ref{EinsteinZ4}) allowing for
132: the contracted Bianchi identities. Here again, by allowing for non-zero values of
133: the extra four-vector $Z_\mu$, one is extending the solution space. But now, as we
134: will see in the following section, one is using all the field equations to evolve
135: the pair $\{ g_{\mu\nu}, Z_\mu \}$: no equation is taken out of the system and, as
136: a result, general covariance is not broken. The initial metric $g_{\mu\nu}$ can be
137: taken to be the one arising from analytical approximations and the initial
138: four-vector $Z_\mu$ can be taken to vanish without any kind of inconsistence: one
139: can even use the evolving values of $Z_\mu$ during the calculation as a good covariant
140: indicator of the quality of the approximation.
141: 
142: Notice that (\ref{WaveZis0}) plays here the role of the subsidiary
143: system. It is adjusted \textit{ab initio}, without any parameter
144: fine-tuning, because light speed is the only characteristic speed
145: in (\ref{WaveZis0}). Constraint deviations, that is non-vanishing
146: values of $Z^\mu$ will then propagate to the boundaries. The fact
147: that our constraints (\ref{Zis0}) are algebraic will greatly
148: simplify the task of providing outgoing boundary conditions that
149: let the constraint deviations get out of the numerical grid
150: \cite{BeA}
151: 
152: Besides these considerations, there are other important
153: theoretical issues that we will address in this work. The first
154: one is a thorough analysis of the hyperbolicity of the evolution
155: system. This is more or less straightforward for the first order
156: version of the system, as discussed in Appendix B, but it is not
157: so well known in the case of the second order version, discussed
158: in Appendix A, where we have used the results of Kreiss and Ortiz
159: \cite{KO01} that recently shed light on this issue, which is
160: crucial to discuss the well-posedness \cite{KL89} of the evolution
161: system (see also \cite{Reu03} for similar results for the BSSN
162: system).
163: 
164: The second theoretical point that we want to stress here is that the Z4 system
165: (\ref{EinsteinZ4}) is not just one more hyperbolic formalism to be added to the
166: long list. As far as it is the only general covariant one, the question arises
167: whether the existing non-covariant hyperbolic formalisms \cite{BM92,BM95,FR94,FR96,SN95,BS99,AY99,Hern00,KST01,ST02}
168: can be recovered from (\ref{EinsteinZ4}) by some ``symmetry breaking" mechanism. We
169: have extended in this sense a previous work \cite{Z3} where the deep relationship
170: between the more widely used hyperbolic formalisms was pointed out. A partial
171: symmetry breaking mechanism is presented in section II for recovering the second
172: order systems \cite{SN95,BS99} and for the first order systems
173: containing additional dynamical fields \cite{BM92,BM95} in section III. These
174: sections are followed by another one containing numerical simulations that have
175: been proposed recently \cite{Mexico} as standard test-beds for Numerical Relativity codes.
176: A more general symmetry breaking mechanism is proposed in Appendix C to recover
177: first order formalisms which do not contain additional dynamical fields
178: \cite{FR94,FR96,AY99,Hern00,KST01,ST02}.
179: 
180: 
181: \section{3+1 Evolution Systems}
182: 
183: 
184: The general-covariant equations (\ref{EinsteinZ4}) can be written
185: in the equivalent 3+1 form \cite{Z4} (Z4~evolution system)
186: \begin{eqnarray}
187: \label{dtgamma}
188:   (\partial_t -{\cal L}_{\beta})~ \gamma_{ij}
189:   &=& - {2\;\alpha}\;K_{ij}
190: \\
191: \label{dtK}
192:    (\partial_t - {\cal L}_{\beta})~K_{ij} &=& -\nabla_i\alpha_j
193:     + \alpha\;   [{}^{(3)}\!R_{ij}
194:     + \nabla_i Z_j+\nabla_j Z_i
195: \nonumber \\
196:     &-& 2K^2_{ij}+(\tr K-2\Theta)\;K_{ij}
197: \nonumber \\
198:     &-& S_{ij}+\frac{1}{2}\,(\tr\; S -\; \tau)\;\gamma_{ij}\;]
199: \\
200: \label{dtTheta}
201: (\partial_t -{\cal L}_{\beta})~\Theta &=& \frac{\alpha}{2}\;
202:  [{}^{(3)}\!R + 2\; \nabla_k Z^k + (\tr K - 2\; \Theta)\;\tr K
203: \nonumber \\
204: &-&  \tr(K^2)  - 2\; Z^k {\alpha}_k/\alpha - 2\tau]
205: \\
206: \label{dtZ}
207:  (\partial_t -{\cal L}_{\beta})~Z_i &=& \alpha\; [\nabla_j\;({K_i}^j
208:   -{\delta_i}^j \tr K) + \partial_i \Theta
209: \nonumber \\
210:   &-&2\; {K_i}^j\; Z_j  -  \Theta\, {{\alpha}_i/ \alpha} - S_i]
211: \end{eqnarray}
212: where we have noted
213: \begin{equation}\label{tauSdef}
214:   \Theta \equiv  \alpha \; Z^0,~
215:   \tau \equiv  8 \pi  \alpha^2\; T^{00},~
216:   S_i \equiv  8 \pi \alpha \; T^0_{\;i},~
217:   S_{ij} \equiv 8 \pi \;T_{ij}.
218: \end{equation}
219: 
220: 
221: In the form (\ref{dtgamma}-\ref{dtZ}), it is evident that the Z4 evolution system
222: consists only of evolution equations. The only constraints (\ref{Zis0}), that can
223: be translated into:
224: \begin{equation}\label{ZThis0}
225:   \Theta~=~0,\qquad Z_i~=~0,
226: \end{equation}
227: are algebraic so that the full set of field equations (\ref{EinsteinZ4}) is
228: actually used during evolution. This is in contrast with the ADM evolution system
229: \cite{ADM62}, which can be recovered from (\ref{dtgamma}-\ref{dtZ}) by imposing
230: (\ref{ZThis0}). The first two equations (\ref{dtgamma},\ref{dtK}) would transform
231: into the well known ADM evolution system, whereas the
232: last two equations (\ref{dtTheta},\ref{dtZ}) would transform into the
233: standard energy and momentum constraints, namely
234: \begin{eqnarray}
235:  {}^{(3)}\!R + \tr^2 K -  {\tr}(K^2)  &=& 2\; \tau
236: \label{constraint1} \\
237:  \nabla_j\;({K_i}^j -{\delta_i}^j trK) &=& S_i
238: \label{constraint2}
239: \end{eqnarray}
240: 
241: In the ``free evolution" ADM approach \cite{Cent80}, both (\ref{constraint1}) and
242: (\ref{constraint2}) were taken out of the evolution system: they were imposed only
243: on the initial data. This was consistent because
244: (\ref{constraint1},\ref{constraint2}) are first integrals of the ADM evolution
245: system, but one can not avoid violations of (\ref{constraint1},\ref{constraint2})
246: due to errors in numerical simulations or approximated initial data, as stated
247: before. As a result, numerical simulations will deal as well with extended solutions.
248: The main difference with the Z4 case, aside from covariance considerations, is that
249: in the Z4 case the quantities $Z_\mu$ describing constraint deviations are included
250: into the evolution system.
251: 
252: One can also ask in this context what happens if, instead of
253: imposing of the full set (\ref{ZThis0}), one imposes the single
254: condition
255: \begin{equation}\label{This0}
256:   \Theta~=~0
257: \end{equation}
258: obtaining a system with only three supplementary dynamical
259: variables $Z_i$ of the kind determined in \cite{Z3} (Z3 system):
260: the one corresponding to the parameter choice
261: \begin{equation}\label{mu2nun0}
262:   \mu~=~2,~~~ \nu~=~n~=~0
263: \end{equation}
264: (we follow the notation of \cite{Z3}).
265: 
266: One can easily understand two of the three conditions (\ref{mu2nun0}),
267: namely
268: \begin{equation}\label{mu2nuisn}
269:   \mu~=~2,~~~ \nu~=~n,
270: \end{equation}
271: because this amounts to the `physical speed' requirement for the
272: degrees of freedom not related to the gauge \cite{Z3} and nothing
273: else can arise from the general covariant equations
274: (\ref{EinsteinZ4}) which are at our starting point. But values of
275: the remaining parameter $n$ other than zero would be very
276: interesting. In particular, the choice
277: \begin{equation}\label{mu2nun43}
278:   \mu~=~2,~~~ \nu~=~n~=~\frac{4}{3}.
279: \end{equation}
280: would lead to the evolution system which is quasiequivalent
281: (equivalent principal parts \cite{Z3}) to the well known BSSN
282: system \cite{SN95,BS99}.
283: 
284: At this point, let us consider the following recombination of the dynamical fields
285: \begin{equation}\label{Ktilde}
286:   \tilde{K}_{ij}~\equiv~K_{ij} - \frac{n}{2}\; \Theta\; \gamma_{ij}
287: \end{equation}
288: so that the Z4 system (\ref{dtgamma}-\ref{dtZ}) can be written in
289: a one-parameter family of equivalent forms just by replacing
290: everywhere
291: \begin{equation}\label{Ktildeinv}
292:   {K_{ij}}~\rightarrow~\tilde{K}_{ij} + \frac{n}{2}\; \Theta\; \gamma_{ij}.
293: \end{equation}
294: This kind of transformations leave invariant the solution space of
295: the system (it is actually the same system expressed in a
296: different set of independent fields). But if the suppression of
297: the $\Theta$ field (\ref{This0}) is made after the replacement
298: (\ref{Ktildeinv}), one gets a one-parameter family of
299: non-equivalent systems (Z3 evolution systems), namely:
300: \begin{eqnarray}
301:   (\partial_t -{\cal L}_{\beta})\; \gamma_{ij}
302:   &=& - {2\;\alpha}\;K_{ij}
303: \label{Z3dtgamma} \\
304: \nonumber
305:    (\partial_t -{\cal L}_{\beta}) K_{ij} &=& -\nabla_i\alpha_j
306:     + \alpha\;   [{}^{(3)}\!R_{ij}
307:     + \nabla_i Z_j + \nabla_j Z_i
308: \nonumber \\
309:     &-& 2K^2_{ij}+\tr K\;K_{ij}
310:  - S_{ij}+\frac{1}{2}(\tr S - \tau)\gamma_{ij}]
311: \nonumber \\
312:  &-& \frac{n}{4} \; \alpha\; [{}^{(3)}\!R + 2\; \nabla\!\! \cdot\! Z
313:  + \tr^2 K - \tr(K^2)
314: \nonumber \\
315:  &&\;\;\;\;\;\;\;\; - 2 (\alpha^{-1} {\alpha_k }) Z^k - 2 \tau]\;\gamma_{ij}
316: \label{Z3dtK} \\
317: \label{Z3dtZ}
318:  (\partial_t -{\cal L}_{\beta}) Z_i &=& \alpha\; [\nabla_j\;({K_i}^j
319:   -{\delta_i}^j~ \tr K) -2 {K_i}^j Z_j - S_i]
320: \nonumber \\
321: \end{eqnarray}
322: where we have suppressed the tilde over $K_{ij}$, allowing for
323: the vanishing of $\Theta$.
324: 
325: The resulting system (\ref{Z3dtgamma}-\ref{Z3dtZ}) is
326: quasiequivalent (equivalent principal parts) to the `system A' in
327: ref. \cite{Z3} verifying the `physical speed' requirement
328: (\ref{mu2nuisn}). As we have already mentioned, it follows that
329: the particular case
330: \begin{equation}\label{nis43}
331:   n ~=~ \frac{4}{3}
332: \end{equation}
333: is quasiequivalent to the BSSN system \cite{SN95,BS99}. The system (\ref{Z3dtgamma}-\ref{nis43}) can be decomposed into trace and trace-free
334: parts
335: \begin{equation}\label{conformal_metric}
336:    e^{4\;\phi} ={\gamma}^{1/3} \;,\quad
337:   {\tilde{\gamma}}_{ij} = e^{-4\;\phi}\;\gamma_{ij}
338: \end{equation}
339: \begin{equation}\label{conformal_curvature}
340:   K = \gamma^{ij}\;K_{ij} \;,\quad
341:   {\tilde{A}}_{ij} = e^{-4\;\phi}\;(K_{ij}-
342:   \frac{1}{3}\; K\;\gamma_{ij})
343: \end{equation}
344: \begin{equation}\label{Gs2}
345:   {\tilde{\Gamma}}_i = -{\tilde{\gamma}}_{ik}\;{{\tilde{\gamma}}^{kj}}_{\;,j}
346:                        + 2\; Z_i
347: \end{equation}
348: to follow the correspondence with BSSN more closely. It must be pointed out,
349: however, that one does not get in this way the original BSSN system: there is
350: actually one difference in the lower order terms (only the principal parts are
351: equivalent). The difference is in the term of the form
352: \begin{equation}\label{Z3AZterm}
353: +\frac{n}{2} \alpha_k\; Z^k\; \gamma_{ij}
354: \end{equation}
355: in the evolution equation (\ref{Z3dtK}), which is missing in the original
356: BSSN system \cite{BS99}. This lower order term is needed for consistency with the general
357: covariant equations (\ref{EinsteinZ4}).
358: 
359: We have seen then how the widely used ADM and BSSN systems can be obtained
360: from the more general Z4 formalism. The equivalence transformation
361: (\ref{Ktilde}) plays the crucial role because suppressing the
362: $\Theta$ field (\ref{This0}) produces a sort of symmetry breaking:
363: different values of the parameter $n$ will lead to evolution
364: systems that can no longer be transformed one into another once
365: the set of dynamical fields is reduced by the disappearance of
366: $\Theta$. It can be regarded as a partial symmetry breaking mechanism for the
367: original equations (\ref{dtgamma}-\ref{dtTheta}). The terms ``partial'' refers to the
368: fact that only the quantity $\Theta$ is suppressed, while the $Z_i$ are kept into the
369: system (\ref{Z3dtgamma}-\ref{Z3dtZ}). A complete symmetry breaking mechanism is
370: discussed in Appendix C.
371: 
372: In section IV, we present the results of some test-bed
373: simulations for the ADM and Z4 systems. We have considered for
374: simplicity only vacuum space-times with the time coordinate
375: conditions
376: \begin{equation}\label{dtAlpha}
377:  (\partial_t -{\cal L}_{\beta})~\ln \alpha = -~\alpha\;[f {\tr} K - \lambda \Theta]
378: \end{equation}
379: which are a further generalization of the one proposed in
380: \cite{Z4}, where
381: \begin{equation}\label{f1lambda2}
382:   f~=~1,~~~ \lambda~=~2.
383: \end{equation}
384: This two-parameter family of coordinate conditions is very
385: interesting from the point of view of Numerical Relativity
386: applications. But it is also interesting from the theoretical
387: point of view, because it provides the opportunity to apply
388: the recent results of Kreiss and Ortiz \cite{KO01} on the hyperbolicity
389: of the ADM system in a wider context. In Appendix A, we will use
390: the same formulation (see ref. \cite{KL89} for more details)
391: to study the hyperbolicity of the Z4 system
392: with the two-parameter family of dynamical gauge conditions
393: (\ref{dtAlpha}).
394: 
395: \section{First Order Systems}
396: 
397: A first order version of the Z4 evolution system
398: (\ref{dtgamma}-\ref{dtZ}) can be obtained in the standard way by
399: considering the first space derivatives
400: \begin{equation}\label{AkDkij}
401:  A_k~\equiv~\alpha_k/\alpha,~~D_{kij}~\equiv~\frac{1}{2}\;\partial_k \gamma_{ij}
402: \end{equation}
403: as independent dynamical quantities with evolution equations
404: given by
405: \begin{equation}\label{dtA}
406:  \partial_t A_k~+~\partial_k [~ \alpha (f\; {\tr}\; K - \lambda \Theta) ~]~=~0
407: \end{equation}
408: \begin{equation}\label{dtD}
409:  \partial_t D_{kij}~+~\partial_k [ \alpha K_{ij} ]~=~0
410: \end{equation}
411: (we will consider in what follows the vanishing shift case for
412: simplicity), so that the full set of dynamical fields can be given
413: by
414: \begin{equation}\label{uvector}
415:  u ~ = ~ \{\alpha,~\gamma_{ij},~ K_{ij},~ A_k,~D_{kij},~\Theta,~Z_k \}
416: \end{equation}
417: (38 independent fields).
418: 
419: Care must be taken when expressing the Ricci tensor
420: ${}^{(3)}\!R_{ij}$ in (\ref{dtK}) in terms of the derivatives
421: of $D_{kij}$, because as far as the constraints
422: (\ref{AkDkij}) are no longer enforced, the identity
423: \begin{equation}\label{dDdD}
424:  \partial_r D_{sij} = \partial_s D_{rij}
425: \end{equation}
426: can not be taken for granted in first order systems. As a
427: consequence of this ordering ambiguity of second derivatives, the
428: principal part of the evolution equation (\ref{dtK}) can be
429: written in a one-parameter family of non-equivalent ways, namely
430: \begin{equation}\label{PrincipalK}
431:  \partial_t K_{ij} ~+~\partial_k~[\alpha~\lambda^k_{ij}]~=~...~
432: \end{equation}
433: \begin{eqnarray}\label{deflambda}
434:  \lambda^k_{ij} &\equiv& -{\Gamma}^k_{~ij}+
435:  \frac{1 - \zeta}{2}\, (D_{ij}^{~~k}+D_{ji}^{~~k}
436:    -\delta^k_i D_{rj}^{~~r}-\delta^k_j D_{ri}^{~~r})
437: \nonumber \\
438:   &+& {\frac{1}{2}}\, \delta^k_i(A_j+{D_{jr}}^r-2Z_j)+
439:   {\frac{1}{2}}\, \delta^k_j(A_i+{D_{ir}}^r-2Z_i)
440: \nonumber \\
441: \end{eqnarray}
442: so that the parameter choice $\zeta = +1$ corresponds
443: to the standard Ricci decomposition
444: \begin{equation}\label{Def3R}
445: {}^{(3)}\!R_{ij}~=~\partial_k\;{\Gamma^k}_{ij}-\partial_i\;{\Gamma^k}_{kj}
446: +{\Gamma^r}_{rk}{\Gamma^k}_{ij}-{\Gamma^k}_{ri}{\Gamma^r}_{kj}
447: \end{equation}
448: whereas the opposite choice $\zeta = -1$ corresponds to the
449: de~Donder-Fock \cite{DeDo21,Fock59} decomposition
450: \begin{eqnarray}\label{Def3dDF}
451: {}^{(3)}\!R_{ij}&=&-\partial_k\;{D^k}_{ij}+\partial_{(i}\;{\Gamma_{j)k}}^{k}
452: - 2 {D_r}^{rk} D_{kij} \nonumber \\
453: &+& 4 {D^{rs}}_i D_{rsj} - {\Gamma_{irs}} {\Gamma_j}^{rs}-{\Gamma_{rij}} {\Gamma^{rk}}_k
454: \end{eqnarray}
455: which is most commonly used in Numerical Relativity codes.  Note
456: that this ambiguity does not affect to the principal part of eq.
457: (\ref{dtTheta}), namely
458: \begin{equation}
459:  \partial_t\Theta + \partial_k [\alpha V^k] = ...
460: \end{equation}
461: where we have noted
462: \begin{equation}\label{defVk}
463:  V_k \equiv {D_{kr}}^r - {D^r}_{rk} -Z_k.
464: \end{equation}
465: 
466: We are now in position to discuss the hyperbolicity of this first
467: order version of the Z4 systems. This is done in a straightforward
468: way in Appendix A.
469: 
470: In order to compare the new first order system with the Bona-Mass\'o ones
471: \cite{BM92,BM95} we could either apply here again the
472: recombination (\ref{Ktilde}) followed by the suppression
473: (\ref{This0}) of the $\Theta$ field or we could take directly a
474: first order version of the Z3 system
475: (\ref{Z3dtgamma}-\ref{Z3dtZ}).
476: Eqs. (\ref{Z3dtgamma},\ref{Z3dtZ},\ref{dtD})  do not change, but eqs.
477: (\ref{dtA},\ref{PrincipalK},\ref{deflambda}) are
478: replaced in this case by
479: \begin{eqnarray}
480: \label{PPZ3dtA}
481:   \partial_t A_k +\partial_k~[\alpha~f~\tr K]&=&0 \\
482: \label{PPZ3dtK}
483:  \partial_t K_{ij} ~+~\partial_k~[\alpha~{\lambda^k}_{ij}]&=& \ldots
484: \end{eqnarray}
485: \begin{eqnarray}\label{PPZ3deflambda}
486:  \lambda^k_{ij} &\equiv& -{\Gamma}^k_{~ij}
487:  -\frac{n}{2}\,V^k\gamma_{ij}
488: \\
489:  &+& \frac{1}{2} \delta^k_i(A_j+{D_{jr}}^r-2Z_j)
490:  + \frac{1}{2} \delta^k_j(A_i+{D_{ir}}^r-2Z_i)
491: \nonumber \\
492:   &+& \frac{1-\zeta}{2} (D_{ij}^{~~k}+D_{ji}^{~~k}
493:    -\delta^k_i D_{rj}^{~~r}-\delta^k_j D_{ri}^{~~r})
494: \nonumber
495: \end{eqnarray}
496: The full Bona-Masso family of evolution equations is recovered for
497: the $\zeta=-1$ case, where (\ref{PPZ3deflambda}) can be written as
498: \begin{eqnarray}\label{PPBMdeflambda}
499:  \lambda^k_{ij} &\equiv& {D}^k_{~ij}-\frac{n}{2} V^k\gamma_{ij}
500: \\
501:  &+&\frac{1}{2} \delta^k_i(A_j-{D_{jr}}^r+2V_j)+
502:  \frac{1}{2} \delta^k_j(A_i-{D_{ir}}^r+2V_i)
503: \nonumber
504: \end{eqnarray}
505: with $V_k$ defined by (\ref{defVk}).
506: 
507: In the following section, we will compare the behavior of both
508: families in numerical simulations. To this end, we will also
509: consider the first order version of the ADM system, which can be
510: obtained from the previous ones just by suppressing the $Z_i$
511: eigenfields.
512: \begin{equation}\label{Z3is0}
513:   Z_i~=~0~.
514: \end{equation}
515: 
516: Before proceeding to the test section, let us just mention, that
517: the same game of recombining the $\Theta$ field with $K_{ij}$
518: (\ref{Ktildeinv}) before suppressing it can also be played with
519: the $Z_k$ fields and $D_{kij}$ in first order systems. As stated before,
520: this will provide a complete symmetry breaking mechanism. We will do
521: that in Appendix C, where we will show how the well known KST system
522: \cite{KST01} can also be recovered in that way from
523: the Z4 framework discussed in this paper.
524: 
525: 
526: \section{Testing Second and First Order Systems}
527: 
528: We will present in this section a couple of numerical experiments
529: which have been suggested very recently \cite{Mexico} as standard
530: test-beds for Numerical Relativity codes.  Our philosophy is that
531: all the tests could be done ``out of the box", using well known
532: numerical methods and the equations that are fully presented here:
533: anyone should be able to reproduce our results without recourse to
534: additional information.
535: 
536: We will use the standard method of lines \cite{MoL} as a finite differencing
537: algorithm, so that space and time discretization will be dealt separately.
538: Space differencing will consist of taking centered
539: discretizations of derivatives in our 3D grid. We use the standard centered
540: stencil for first derivatives and we make sure that second derivatives,
541: when needed, are coded also as centered derivatives of these first
542: derivatives, even if it takes up to five point along every axis.In order to
543: avoid boundary effects, the grid has the topology of a three-torus, with
544: periodic boundaries along every axis. The time evolution will be
545: dealt with a third order Runge-Kutta algorithm. The time step ${\rm d}t$ is kept
546: small enough to avoid an excess of numerical dissipation that could distort
547: our results in long runs.
548: 
549: 
550: \subsection{Robust stability test}
551: 
552: Let us consider a small perturbation of Minkowski space-time which
553: is generated by taking random initial data for every dynamical
554: field in the system. The level of the random noise must be small
555: enough to make sure that we will keep in the linear regime even
556: for a thousand of crossing times (the time that a light ray will
557: take to cross the longest way along the numerical domain). This is
558: in keeping with the theoretical framework of Appendix A, where
559: only linear perturbations around the Minkowski metric are
560: considered.
561: 
562: We have plotted in Fig.~\ref{plot1} our results for the standard harmonic
563: case (\ref{f1lambda2}). We see the expected polynomial (linear in this case)
564: growth \cite{KL89} of the weakly hyperbolic ADM system.
565: Notice that modifications of the lower order terms (the ones not
566: contributing to the principal part) could lead to catastrophic
567: exponential growth, revealing an ill-posed evolution system
568: \cite{KL89}. In this paper, however, we will limit ourselves to discussing the linear
569: regime as an hyperbolicity test for the principal part of the system. In this sense,
570: the linear growth of the ADM plots in Fig.~\ref{plot1} confirms
571: the weakly hyperbolic character of the ADM system.
572: 
573: The Z4 system shows instead the no-growth behavior, independent
574: of the time resolution (see Figs.~\ref{plot1},~\ref{plot1bis}), which one would expect from
575: a strongly hyperbolic system. The same qualitative behavior is shown by the
576: corresponding Z3 systems in (\ref{Z3dtgamma}-\ref{Z3dtZ}),
577: including the one with $n=4/3$ that is quasiequivalent to the BSSN
578: system.
579: 
580: \begin{figure}[t]
581: \begin{center}
582: \epsfxsize=8cm
583: \epsfbox{fig1.ps}
584: %\hspace{8mm}
585: %\epsfxsize=8cm
586: \end{center}
587: \caption{\label{plot1} The maximum of (the absolute value of) $trK$ is
588: plotted against the number of crossing times in a logarithmic scale. The
589: initial level of random noise remains constant during the evolution in the
590: case of strongly hyperbolic systems (only the second order Z4 system is shown
591: here for clarity). In the case of weakly hyperbolic systems, like the ADM
592: second order system ADM-2 or its first order version ADM-1, a linear growth
593: is detected up to the point where the codes crash. Notice that the second
594: order version is more robust, an order of magnitude, than the first one.
595: The simulations are made with 50 grid points with ${\rm d}t=0.03\;{\rm d}x$.}
596: \end{figure}
597: 
598: \begin{figure}[t]
599: \begin{center}
600: \epsfxsize=8cm
601: %\hspace{8mm}
602: %\epsfxsize=8cm
603: \epsfbox{fig1bis.ps}
604: \end{center}
605: \caption{\label{plot1bis} Same as Fig.~\ref{plot1}, but with less time resolution
606: (${\rm d}t=0.06\;{\rm d}x$ with the same ${\rm d}x$). There is a slight amount of
607: dissipation that delays the crashing of the ADM codes; this is especially visible
608: for the second order version ADM-2, which keeps being more robust than its first order
609: counterpart ADM-1. The behaviour of the Z4 code keeps unaffected.}
610: \end{figure}
611: 
612: \begin{figure}[b]
613: \caption{\label{plot3} Array of results of numerical experiments
614: in the gauge parameter plane ($f$,$\lambda$), by using the Z4
615: system. A triangle stands for linear growth of noise (weak
616: hyperbolicity), whereas a cross stands for a constant noise level
617: (strong hyperbolicity). This is consistent with the strong
618: hyperbolicity requirements predicted in Appendix B: either $f =1$
619: and $\lambda=2$, or $f \neq 1$ and $f> 0$.}
620: \begin{center}
621: %\epsfxsize=11cm % will enlarge or reduce the postscript figures based on the xsize
622: \epsfbox{fig3.ps}
623: \end{center}
624: \end{figure}
625: 
626: \begin{figure}[t]
627: \begin{center}
628: \epsfxsize=8cm
629: %\hspace{8mm}
630: %\epsfxsize=8cm
631: \epsfbox{fig2.ps}
632: \end{center}
633: \caption{\label{plot2} Same as Fig.~\ref{plot1}, but using the
634: second order ICN method to evolve in time instead of a third order
635: Runge-Kutta algorithm. Numerical dissipation is severely
636: distorting the plots, by masking the linear growth in the weakly
637: hyperbolic case and dramatically reducing the initial noise level
638: in the strongly hyperbolic case. Notice than both ${\rm d}t$ and
639: ${\rm d}x$ are the same as in Fig.~\ref{plot1} and we are using
640: also the same space discretization algorithm: only the time
641: evolution method has changed.}
642: \end{figure}
643: 
644: 
645: We also show in Fig.~\ref{plot2} the same results, but distorted by using
646: too much numerical dissipation: the time evolution here is dealt
647: with the second order ICN method rather than the third order
648: Runge-Kutta of Fig.~\ref{plot1}. After hundreds of crossing times, the
649: numerical dissipation manages to curve the linear growing of ADM
650: and the noise level goes down in the Z4 case. This is just a
651: numerical artifact, because in the linear regime there is no
652: physical damping mechanism for strongly hyperbolic systems in a
653: three-torus, where periodic boundary conditions do not allow
654: propagation outside the domain. This is why we will use here third
655: order Runge-Kutta instead of the ICN method proposed in
656: \cite{Mexico}.
657: 
658: In Fig.~\ref{plot3}, we explore parameter space in the ($f$,$\lambda$)
659: plane. If we interpret the constant behavior in Fig.~\ref{plot1} as
660: revealing a strongly hyperbolic system and the polynomial growth
661: (linear in this case) in Fig.~\ref{plot1} as revealing a weakly hyperbolic system, the
662: results of our numerical experiment fully agree with the
663: theoretical results presented in Appendix A.
664: 
665: \subsection{Gowdy waves}
666: 
667: In order to test the strong field regime, let us consider now the Gowdy solution
668: \cite{Gowdy71}, which describes a space-time containing plane polarized gravitational
669: waves (see also \cite{Berger} for an excellent review of these space-times as
670: cosmological models). The line element can be written as
671: \begin{equation}\label{gowdy_line}
672:   {\rm d}s^2 = t^{-1/2}\, e^{Q/2}\,(-{\rm d}t^2 + {\rm d}z^2)
673:   + t\,(e^P\, {\rm d}x^2 + e^{-P}\, {\rm d}y^2)
674: \end{equation}
675: where the quantities $Q$ and $P$ are functions of $t$ and $z$ only
676: and periodic in $z$, so that (\ref{gowdy_line}) is well suited for
677: finite difference numerical grids with periodic boundary
678: conditions along every axis. Following \cite{Mexico}, we will
679: choose the particular case
680: \begin{eqnarray}
681: \label{function_P}
682:   P &=& J_0 (2 \pi t)\; \cos(2 \pi z)
683: \\
684: \label{fucntion_L}
685:   Q &=&  \pi J_0 (2 \pi) J_1 (2 \pi)
686:    -2 \pi t J_0 (2 \pi t) J_1 (2 \pi t) \cos^2(2 \pi z)
687: \nonumber \\
688:   &+& 2 \pi^2 t^2 [{J_0}^2 (2 \pi t)  + {J_1}^2 (2 \pi t)
689:   - {J_0}^2 (2 \pi)  - {J_1}^2 (2 \pi)]
690: \nonumber \\
691: \end{eqnarray}
692: so that it is clear that the lapse function
693: \begin{equation}\label{gowdy_lapse}
694:   \alpha = t^{-1/4}\; e^{Q/4}
695: \end{equation}
696: is constant everywhere at any time $t_0$ at which $J_0(2\pi t_0)$ vanishes. In \cite{Mexico}
697: the initial slice $t=t_0$ was chosen for the simulation of the collapse,
698: where $2 \pi t_0$ is the 20-th root of the Bessel function $J_0$, i.e. $t_0\; \simeq \;9.88$.
699: 
700: \begin{figure}[t]
701: \begin{center}
702: \epsfxsize=8cm % will enlarge or reduce the postscript figures based on the xsize
703: \epsfbox{fig-lapse.ps}
704: \end{center}
705: \caption{\label{plot4} Time evolution of (the maximum value of) the lapse function
706: $\alpha$ in a collapsing Gowdy space-time (harmonic slicing). Notice that the harmonic
707: time coordinate $\tau$ is not the proper time and it does not coincide with the
708: number of crossing times, due to the collapse of the lapse, which is visible here, by a
709: $10^{-9}$ factor.}
710: \end{figure}
711: 
712: Let us now perform the following time coordinate transformation
713: \begin{equation}\label{gowdy_time}
714:   t~=~t_0\;e^{-\tau / \tau_0}, ~~ \tau_0~=~t_0^{3/4} e^{Q(t_0)/4}~\simeq~472\;,
715: \end{equation}
716: so that the expanding line element (\ref{gowdy_line}) is seen in the new time
717: coordinate $\tau$ as collapsing towards the $t=0$ singularity, which is approached
718: only in the limit $\tau\rightarrow\infty$. This ``singularity avoidance" property of the $\tau$
719: coordinate is not surprising if one realizes that the resulting slicing by $\tau=constant$
720: surfaces is harmonic \cite{BM83}.
721: 
722: \begin{figure}[b]
723: \caption{\label{plot5} The quantity $\Theta$ is plotted as an
724: indicator of the accumulated error of the simulations for the ADM,
725: Z4 and Z3-BSSN second order forms. Even in this logarithmic scale,
726: it can be clearly seen how the Z4 and Z3-BSSN codes perform much
727: better than the ADM one: error differs by one order of magnitude
728: at $\tau \simeq 1000$. The Z3-BSSN code gets closer to the Z4 one
729: in the oscillatory phase (up to $\tau \simeq 2000$).}
730: \begin{center}
731: \epsfxsize=8cm % will enlarge or reduce the postscript figures based on the xsize
732: %\epsfbox{gow-theta.eps}
733: \epsfbox{fig-g2.ps}
734: \end{center}
735: \end{figure}
736: 
737: This means that we can launch our simulations starting with a
738: constant lapse $\alpha_0=1$ at $\tau=0$ ($t=t_0$) with the gauge
739: parameter choice $f=1$ (which means also $\lambda=2$ in the Z4
740: case). Notice that the harmonic time coordinate $\tau$ is not the
741: proper time and it does not coincide with the number of crossing
742: times, due to the collapse of the lapse. Remember also that the
743: local value of light speed (proper distance over coordinate time)
744: is $\pm \alpha$. Even though in our plots $\tau$ goes up to
745: $10000$, the light ray manages to cross the domain in the
746: $z$-direction only $t_0 \simeq 9.88$ times, as it follows from the
747: original form (\ref{gowdy_line}) of the line element.
748: 
749: 
750: \begin{figure}[t]
751: \begin{center}
752: \epsfxsize=8cm % will enlarge or reduce the postscript figures based on the xsize
753: %\epsfbox{gow-theta1st.eps}
754: \epsfbox{fig8.ps}
755: \end{center}
756: \caption{\label{plot6bis} Convergence test for the Z4 code. The three lines
757: correspond to $50$, $100$ and $200$ grid points. The quantity $\Theta$ itself is a
758: direct measure of the error. In that logarithmic scale differences of $log\, 4$
759: correspond to dividing by four the error when doubling the resolution. This second
760: order convergence rate is clearly shown in the figure.}
761: \end{figure}
762: 
763: \begin{figure}[b]
764: \caption{\label{plot6} Same as Fig.~\ref{plot5}, but with the
765: first order versions of both the ADM and Z4 codes, which behave in
766: the same way as their second order counterparts. The Z3-BM code
767: gets here closer the ADM one in the oscillatory phase (up to $\tau
768: \simeq 2000$), in contrast with the behavior of the Z3-BSSN code
769: in Fig.~\ref{plot5}.}
770: \begin{center}
771: \epsfxsize=8cm % will enlarge or reduce the postscript figures based on the xsize
772: %\epsfbox{gow-theta1st.eps}
773: \epsfbox{fig-g1.ps}
774: \end{center}
775: \end{figure}
776: 
777: We plot in Fig.~\ref{plot4} the maximum values of the lapse
778: function as time goes on, measured in terms of the harmonic time
779: coordinate $\tau$. Notice the huge magnitude of the dynamical
780: space we are covering, as $\alpha$ goes down (collapse of the
781: lapse) by the factor of one billion during the simulation. This is
782: a  real challenge for numerical codes and all of them are doing
783: quite well until $\tau=1000$. The behavior at later times is
784: dominated by the lower order terms: coordinate light speed ($\pm
785: \alpha$) is so small that the dynamics of the principal part is
786: frozen and care must be taken to avoid too big time steps. We have
787: not seen any of the codes crashing even in very long simulations
788: (up to $\tau = 60.000$) when the size of the time step is kept
789: under control.
790: 
791: In Fig.~\ref{plot5}, we use the quantity $\Theta$, as computed
792: from (\ref{dtTheta}) to monitor the quality of the simulation both
793: in the Z4 case (\ref{dtgamma}-\ref{dtZ}), where $\Theta$ is a
794: dynamical field, and in the other two second order cases (ADM and
795: Z3-BSSN), where $\Theta$ is no longer a dynamical quantity but can
796: still be used as a good measure of the error in the simulation.
797: This makes easier to perform convergence tests (second order convergence
798: is shown in Fig.~\ref{plot6bis}). Notice that our Z3-BSSN code performs
799: here much better than the original BSSN one \cite{BS99}, as reported in
800: \cite{Mexico}. This is mainly due to the fact that we are not using here the conformal
801: decomposition (\ref{conformal_metric}-\ref{Gs2}) which is at odds
802: with the structure of the line element (\ref{gowdy_line}). This is
803: why we talk here about Z3-BSSN (\ref{Z3dtgamma}-\ref{nis43})
804: instead of simply BSSN.
805: 
806: \begin{figure}[b]
807: \caption{\label{plot7} Same as Fig.~\ref{plot5}, but with
808: different values of the parameter $n$ arising from the symmetry
809: breaking mechanism in the Z3 codes. The differences show up in the
810: collapse final phase (starting at $\tau \simeq 2000$). Notice that
811: the value $n=4/3$ corresponds to the Z3-BSSN case in
812: Fig.~\ref{plot5}, whereas the case $n=1$ corresponds to (the
813: second order version of) the Z3-BM case in Fig.~\ref{plot6}. First
814: order versions (not shown) behave in the same way as their second
815: order counterparts shown here}
816: \begin{center}
817: \epsfxsize=8.38cm % will enlarge or reduce the postscript figures based on the xsize
818: \epsfbox{fig-nn.ps}
819: \end{center}
820: \end{figure}
821: 
822: The same kind of comparison is made in Fig.~\ref{plot6} for the
823: first order version of the ADM and Z4 codes, which show the same
824: behavior than their second order counterparts in Fig.~\ref{plot5}.
825: The third plot corresponds to the Z3 version of the Bona-Mass\'o
826: code, which can be obtained from
827: (\ref{PPZ3dtA}-\ref{PPZ3deflambda}), with the parameter choice
828: $\zeta = -1$, $n=+1$. Notice that in the oscillatory phase (up to
829: $\tau \simeq 2000$) the Z3-BM code tends to behave like the ADM
830: one, whereas in Fig.~\ref{plot5} the Z3-BSSN code tends to behave
831: more like the Z4 one.
832: 
833: Notice that the Z3 versions, when compared with both ADM and Z4,
834: show a different behavior after the oscillatory phase: $\theta$
835: grows at a much lower rate or even starts going down. This
836: behavior is due to the extra terms that appear in the evolution
837: equations for $K_{ij}$ after the symmetry breaking, which can be
838: controlled by the parameter $n$ introduced in (\ref{Ktildeinv}).
839: This point is clearly shown in Fig.~\ref{plot7}, where different
840: choices of $n$ produce different behavior in the final collapse
841: phase (starting at $\tau \simeq 2000$). This shows the relevance
842: of the recombination between the dynamical fields in numerical
843: applications, as pointed out in \cite{KST01}. Further details and other
844: numerical tests can be found in the webpage at \textit{http://stat.uib.es}.
845: 
846: 
847: \renewcommand{\theequation}{A.\arabic{equation}}
848: \setcounter{equation}{0}
849: \section*{Appendix A: Hyperbolicity of the Second Order Z4 System}
850: 
851: Let us consider the linearized version of the Z4 system
852: (\ref{dtgamma}-\ref{dtZ}) around Minkowski space-time in order to
853: study the propagation of a plane wave in that background:
854: \begin{eqnarray}
855:   \gamma_{ij} &=& \delta_{ij} + 2\; e^{i\;
856:   \omega \cdot x}\; {\hat{\gamma}}_{ij} (\omega,t)
857: \label{pert_gamma} \\
858:   \alpha &=& 1 +  e^{i\;\omega \cdot x}\; \hat{\alpha}
859: (\omega,t)
860: \label{pert_alpha} \\
861:   K_{ij} &=& i\; \omega\; e^{i\;\omega \cdot x}\;
862: {\hat{K}}_{ij} (\omega,t)
863: \label{pert_K} \\
864:   \Theta &=& i\; \omega\; e^{i\;\omega \cdot x}\;
865: {\hat{\Theta}} (\omega,t)
866: \label{pert_Theta} \\
867:   Z_{k} &=& i\; \omega\; e^{i\;\omega \cdot x}\;
868: {\hat{Z}}_{k} (\omega,t)
869: \label{pert_Z}
870: \end{eqnarray}
871: where we will take for simplicity $\beta^i = 0$ and
872: \begin{equation}\label{shortcuts}
873:   \omega_k = \omega\;n_k \:,\qquad \delta^{ij}\; n_i\; n_j = 1
874: \end{equation}
875: 
876: The Z4 system reads then
877: \begin{eqnarray}
878: \partial_t {\hat{\gamma}}_{ij} &=& - i\; \omega\; {\hat{K}}_{ij}
879: \label{A_dtgamma} \\
880: \partial_t {\hat{\alpha}} &=& - i\; \omega\; [f\;tr{\hat{K}} -
881: \lambda\;{\hat{\Theta}}]
882: \label{A_dtalpha} \\
883: \partial_t {\hat{\Theta}} &=& - i\; \omega\; [tr{\hat{\gamma}} -
884: {\hat{\gamma}}^{nn}-{\hat{Z}}^n]
885: \label{A_dtTheta} \\
886: \partial_t {\hat{Z}}_k &=& - i\; \omega\; [n_k\; (tr{\hat{K}} -
887: {\hat{\Theta}}) - {{\hat{K}}_k}^n]
888: \label{A_dtZ} \\
889: \partial_t {\hat{K}}_{ij} &=& - i\; \omega\; {\hat{\lambda}}_{ij}
890: \label{A_dtK}
891: \end{eqnarray}
892: where we have noted
893: \begin{eqnarray}\label{A_deflambda}
894: {\hat{\lambda}}_{ij} &\equiv& {\hat{\gamma}}_{ij} +
895: n_i\;n_j\;({\hat{\alpha}} + tr{\hat{\gamma}})
896: \nonumber \\
897: &-& n_i\;({{\hat{\gamma}}_j}^n + {\hat{Z}}_j) -
898: n_j\;({{\hat{\gamma}}_i}^n + {\hat{Z}}_i)
899: \end{eqnarray}
900: and where the symbol $n$ replacing an index means the contraction
901: with $n_i$. It can be also expressed in matrix form, namely
902: \begin{eqnarray}\label{2Z4-linear-matrix}
903:   {\hat{u}} &=& ({\hat{\alpha}},{\hat{\gamma}}_{ij},{\hat{K}}_{ij},
904:   {\hat{\Theta}},{\hat{Z}}_k) \\
905:   \partial_t {\hat{u}} &=& - i\; \omega\; {\bf A}\; {\hat{u}}
906: \end{eqnarray}
907: 
908: The spectral analysis of the characteristic matrix ${\bf A}$ provides
909: the following list of eigenvalues and eigenfields
910: 
911: \renewcommand{\labelenumi}{\alph{enumi})}
912: \begin{enumerate}
913: \item Standing eigenfields (zero characteristic speed)
914: \begin{equation}\label{A_EF0}
915:   {\hat{\alpha}} - f\/ \tr{\hat{\gamma}} + \lambda (\tr{\hat{\gamma}} -
916:   {\hat{\gamma}}^{nn} - {\hat{Z}}^n) \,,\quad
917:   {{\hat{\gamma}}^{n}}_{\perp} + {\hat{Z}}_{\perp}
918: \end{equation}
919: where the symbol $\perp$ replacing an index means the projection orthogonal
920: to $n_i$.
921: \item Light-cone eigenfields (characteristic speed $\pm 1$)
922: \begin{eqnarray}
923:  && {\hat{K}}_{\perp \perp} \pm {\hat{\gamma}}_{\perp \perp}
924: \label{A_EFLtt} \\
925:  && {{\hat{K}}_{\perp}}^n \pm {\hat{Z}}_{\perp}
926: \label{A_EFLnt} \\
927:  && {\hat{\Theta}} \pm [tr{\hat{\gamma}} - {\hat{\gamma}}^{nn} -
928:  {\hat{Z}}^n ] \label{A_EFLen}\;.
929: \end{eqnarray}
930: Notice that (\ref{A_EFLtt}-\ref{A_EFLen}) can
931: be seen as the components of a tensor:
932: \begin{eqnarray}\label{A_EFLij}
933:   {\hat{L}}_{ij}^{\pm} &\equiv& [{\hat{K}}_{ij} - (tr{\hat{K}}-2{\hat{\Theta}})\;
934:   n_i\;n_j]
935: \nonumber \\
936:   &\pm& [{\hat{\lambda}}_{ij} - {\hat{\alpha}}\;n_i\;n_j]
937: \end{eqnarray}
938: 
939: \item Gauge eigenfields (characteristic speed $\pm \sqrt{f}$)
940: \begin{eqnarray}\label{A_EFG}
941:   {\hat{G}}^{\pm} &\equiv& \sqrt{f} \left[\tr{\hat{K}} + \frac{2-\lambda}{f-1}\;{\hat{\Theta}}\right]
942: \nonumber \\
943:    &\pm& \left[{\hat{\alpha}} + \frac{2\;f-\lambda}{f-1}(\tr{\hat{\gamma}}
944:     - {\hat{\gamma}}^{nn} - {\hat{Z}}^n )\right]
945: \end{eqnarray}
946: \end{enumerate}
947: 
948: From (\ref{A_EF0}-\ref{A_EFG}) we can easily conclude \cite{KO01,KL89}
949: \renewcommand{\labelenumi}{\roman{enumi})}
950: \begin{enumerate}
951: \item All the characteristic speeds are real (weak hyperbolicity at least) if and
952: only if $f\geq 0$.
953: 
954: \item In the case $f=0$, the two components of the gauge pair
955: (\ref{A_EFG}) are not independent, so that the total number of
956: independent eigenfields is 16 instead of 17 required for strong
957: hyperbolicity.
958: 
959: \item The case $f=1$ (harmonic case) is special:
960: \begin{itemize}
961: \item if $\lambda \ne 2$, then the gauge pair (\ref{A_EFG}), which
962: can be previously rescaled by a $(f-1)$ factor, is equivalent to
963: (\ref{A_EFLen}), so that one has only 15 independent eigenfields
964: \item if $\lambda=2$, then the quotient $\frac{2-\lambda}{f-1}$
965: can take any value, reflecting the degeneracy of the gauge and
966: light cone eigenfields. One can then recover the full set of 17
967: independent eigenfields (strong hyperbolicity).
968: \end{itemize}
969: \item In all the remaining cases ($f>0$, $f \ne 1$), the system is
970: strongly hyperbolic, as we can recover the full set of 17
971: independent eigenfields.
972: \end{enumerate}
973: 
974: \renewcommand{\theequation}{B.\arabic{equation}}
975: \setcounter{equation}{0}
976: \section*{Appendix B: Hyperbolicity of the First Order Z4 System}
977: 
978: The principal part of the first order Z4 evolution system
979: (\ref{dtgamma}-\ref{dtZ},\ref{dtAlpha},\ref{dtA}-\ref{dtD})
980: can be written as (vanishing shift case):
981: \begin{eqnarray}
982: \label{B_dtgamma}
983: &\!\!\!\!\!& \partial_t \gamma_{ij} = \ldots \;,\qquad  \partial_t \alpha = \ldots
984: \\
985: \label{B_dtTheta}
986: &\!\!\!\!\!& \partial_t \Theta +\partial_k~[\alpha~V^k] = \ldots
987: \\
988: \label{B_dtZ} &\!\!\!\!\!& \partial_t Z_i +
989: \partial_k~[\alpha~(\delta^k_i(\tr K-\Theta)-K^k_{~i})] =\ldots
990: \\
991: \label{B_dtA}
992: &\!\!\!\!\!& \partial_t A_k +\partial_k~[\alpha~(f\/ \tr K - \lambda \Theta)]=\ldots
993: \\
994: \label{B_dtD}
995: &\!\!\!\!\!& \partial_t D_{kij} +\partial_k~[\alpha~K_{ij}] =\ldots
996: \\
997: \label{B_dtK}
998: &\!\!\!\!\!& \partial_t K_{ij} +\partial_k~[\alpha~\lambda^k_{ij}] =\ldots
999: \end{eqnarray}
1000: where
1001: \begin{eqnarray}\label{B_deflambda}
1002:  \lambda^k_{ij} &=& {D^k}_{ij}
1003:    -{\frac{1+\zeta}{2}} (D_{ij}^{~~k}+D_{ji}^{~~k}
1004:    -\delta^k_i D_{rj}^{~~r}-\delta^k_j D_{ri}^{~~r})
1005: \nonumber \\
1006:  &+& \frac{1}{2}\, \delta^k_i(A_j-{D_{jr}}^r+2V_j)
1007: \nonumber \\
1008:  &+& \frac{1}{2}\, \delta^k_j(A_i-{D_{ir}}^r+2V_i)
1009: \\
1010: \label{B_defV}
1011:   V_k &\equiv& {D_{kr}}^r-{D_{rk}}^{r}-Z_k.
1012: \end{eqnarray}
1013: 
1014: Now, if we consider the propagation of perturbations with
1015: wavefront surfaces given by the unit (normal) vector $n_i$, we can
1016: express (\ref{B_dtgamma}-\ref{B_dtK}) in matrix form
1017: \begin{equation}\label{B_dtu}
1018:   \alpha^{-1} \partial_t~u ~+~ {\bf A}(u) ~n^k~\partial_k u = ...~,
1019: \end{equation}
1020: where
1021: \begin{equation}\label{B_uvector}
1022:  u ~ = ~ \{\alpha,~\gamma_{ij},~ K_{ij},~ A_k,~D_{kij},~\Theta,~Z_k \}~.
1023: \end{equation}
1024: (notice that derivatives tangent to the wavefront surface play no
1025: role here).
1026: 
1027: A straightforward analysis of the characteristic matrix ${\bf A}(u)$
1028: provides the following list of eigenfields:
1029: 
1030: \renewcommand{\labelenumi}{\alph{enumi})}
1031: \begin{enumerate}
1032: \item Standing eigenfields (zero characteristic speed)
1033: \begin{equation}\label{B_EF0}
1034:  \alpha,~ \gamma_{ij},~A_\perp,~D_{\perp ij},~A_k-f D_k+\lambda V_k
1035: \end{equation}
1036: (24 independent fields), where the symbol $\perp$ replacing an
1037: index means the projection orthogonal to $n_i$:
1038: \begin{equation}\label{B_DAij}
1039:  D_{\perp ij} \equiv D_{kij} - n_k n^r D_{rij}.
1040: \end{equation}
1041: 
1042: \item Light-cone eigenfields (characteristic speed $\pm 1$)
1043: \begin{eqnarray}\label{B_EFLij}
1044:  {L^{\pm}}_{ij} &\equiv& [K_{ij}-n_i n_j \;{\tr}K ]
1045: \nonumber \\
1046:   &\pm& [{\lambda^n}_{ij} - n_i n_j~{\tr}\;\lambda^n]
1047: \\
1048: \label{B_EFL}
1049:   L^{\pm} &\equiv& \theta \pm V^n
1050: \end{eqnarray}
1051: (12 independent fields), where the symbol $n$ replacing the index
1052: means the contraction with $n_i$
1053: \begin{equation}
1054:  \lambda^n_{ij}~\equiv~n_k~\lambda^n_{ij}.
1055: \end{equation}
1056: 
1057: \item Gauge eigenfields (characteristic speed $\pm \sqrt{f}$)
1058: \begin{eqnarray}\label{B_EFG}
1059:  G^{\pm} &\equiv& \sqrt{f} \left[{\tr}~K + \frac{2-\lambda}{f-1}\,\Theta \right]
1060: \nonumber \\
1061:   &\pm& \left[A^n + \frac{2f-\lambda}{f-1}\,V^n \right]
1062: \end{eqnarray}
1063: \end{enumerate}
1064: 
1065: 
1066: From (\ref{B_EF0}-\ref{B_EFG}) we can easily conclude that
1067: 
1068: \renewcommand{\labelenumi}{\roman{enumi})}
1069: \begin{enumerate}
1070: \item All the characteristic speeds are real (weak hyperbolicity
1071: at least) if and only if $f\geq 0$.
1072: 
1073: \item In the case $f=0$, the two components of the pair
1074: (\ref{B_EFG}) are not independent, so that the total number of
1075: independent eigenfields is 37 instead of 38 required for strong
1076: hyperbolicity.
1077: 
1078: \item The case f=1 (harmonic case) is special:
1079: \begin{itemize}
1080: \item if $\lambda \ne 2$, then the pair of fields (\ref{B_EFG}) is
1081: the same as (\ref{B_EFL}), so that one has only 36 independent
1082: eigenfields \item if $\lambda=2$, then the quotient
1083: $\frac{2-\lambda}{f-1}$ can take any value due to the degeneracy
1084: of the gauge and light eigenfields. One can then recover the full
1085: set of 38 independent eigenfields (strong hyperbolicity).
1086: \end{itemize}
1087: \item The first order Z4 system described by
1088: (\ref{B_dtgamma}-\ref{B_dtK}) is strongly hyperbolic in all the
1089: remaining cases ($f>0$, $f \ne 1$).
1090: \end{enumerate}
1091: 
1092: Notice also that the special (harmonic) case $f=1$, $\lambda=2$
1093: has been shown in \cite{Z4} to be symmetric hyperbolic for the
1094: parameter choice $\zeta=-1$. The corresponding energy function
1095: can be written as
1096: \begin{eqnarray}\label{B_Energy}
1097:   E &\equiv&
1098:  K^{ij}K_{ij} + \lambda^{kij}\lambda_{kij} +
1099:   (\tr{K} - 2\Theta)^2 + A^k A_k
1100: \nonumber \\
1101:   &+& (A^k - {D^{kr}}_r + 2 V^k)(A_k - {D_{kr}}^r + 2 V_k)
1102: \end{eqnarray}
1103: but notice that this expression is far from being unique. For
1104: instance, allowing for (\ref{B_EF0}), the last term in
1105: (\ref{B_Energy}) could appear with any arbitrary factor.
1106: 
1107: \renewcommand{\theequation}{C.\arabic{equation}}
1108: \setcounter{equation}{0}
1109: \section*{Appendix C: Recovering the KST systems}
1110: 
1111: Let us start with the first order Z4 evolution system
1112: (\ref{dtgamma}-\ref{dtZ},\ref{dtAlpha},\ref{dtA}-\ref{dtD}) where
1113: the principal part is given by (\ref{B_dtgamma}-\ref{B_defV}). Now
1114: let us follow the two step `symmetry breaking' process, namely
1115: 
1116: \begin{enumerate}
1117: \item Recombine the dynamical fields $K_{ij}$, $D_{kij}$ with
1118: $\Theta$ and $Z_i$ in a linear way,
1119: \begin{eqnarray}
1120: \label{C_Ktilde}
1121:   \tilde{K}_{ij}&=&K_{ij} - {\frac{n}{2}}\; \Theta\; \gamma_{ij}~,
1122: \\
1123: \label{C_dtilde}
1124:   d_{kij}&=& 2 D_{kij} + \eta\; \gamma_{k(i} Z_{j)} + \chi\; Z_k \gamma_{ij}~,
1125: \end{eqnarray}
1126: where we have used the notation of Ref. \cite{KST01}, replacing
1127: only their parameter $\gamma$ by $-n/2$ for consistency. Notice that
1128: (\ref{C_Ktilde}-\ref{C_dtilde}) is generic in the sense that it is the most
1129: general linear combination that preserves the tensor character of the dynamical fields
1130: under linear coordinate transformations.
1131: 
1132: \item Suppress both $\theta$ and $Z_i$ as dynamical fields, namely
1133: \begin{equation}\label{C_ZThis0}
1134:   \Theta~=~0,\qquad Z_i~=~0~.
1135: \end{equation}
1136: \end{enumerate}
1137: 
1138: 
1139: In that way, the principal part (\ref{B_dtgamma}-\ref{B_defV})
1140: becomes
1141: \begin{eqnarray}\label{C_dtgamma}
1142:  \partial_t \gamma_{ij} &=& \ldots \;, \qquad \partial_t \alpha = \ldots
1143: \\
1144: \label{C_dtA}
1145:  \partial_t A_k &+&\partial_k~[\alpha~f\;\tr\; \tilde{K}\; ]~=~0
1146: \\
1147: \label{C_dtd}
1148:    \partial_t d_{kij} &+& \partial_r [
1149:     \alpha \{2\; \delta^r_k\; \tilde{K}_{ij}
1150:     -\chi\; ({\tilde{K}_{k}}^r-\delta^r_k\; \tr \tilde{K}) \gamma_{ij}
1151: \nonumber \\
1152:    &+& \eta\; \gamma_{k(i}({{\tilde{K}}^r}_{~j)}   - {\delta^r}_{j)} \tr \tilde{K})\}] =
1153:    \ldots
1154: \\
1155: \label{C_dtK}
1156:  \partial_t \tilde{K}_{ij} &+&\partial_k~[\alpha~\lambda^k_{ij}]=\ldots
1157: \end{eqnarray}
1158: \begin{eqnarray}\label{C_deflambda}
1159:    2\; {\lambda^k}_{ij} &=& {d^k}_{ij} - \frac{n}{4}({d_{kr}}^r-{{d_r}^{rk}})\gamma_{ij}
1160: \nonumber \\
1161:    &+& \frac{1+\zeta}{2}({d_{ij}}^k + {d_{ji}}^k )
1162:    - \frac{1-\zeta}{2}( \delta_i^k\, {d_{rj}}^r+ \delta_j^k\, {d_{ri}}^r)
1163: \nonumber \\
1164:    &+& \delta^k_j\, (A_i+\frac{1}{2} {d_{ir}}^r)+ \delta^k_i\, (A_j+\frac{1}{2} {d_{jr}}^r)
1165: \end{eqnarray}
1166: 
1167: 
1168: for the reduced set of variables
1169: \begin{equation}\label{C_uvector}
1170:  u ~ = ~ \{\alpha,~\gamma_{ij},~ \tilde{K}_{ij},~ A_k,~d_{kij} \}.
1171: \end{equation}
1172: 
1173: This provides a ``dynamical lapse" version \cite{ST02} of the KST
1174: evolution systems. In order to recover the original ``densitized
1175: lapse" version, one must in addition integrate explicitly  the
1176: dynamical relationship (\ref{dtAlpha}) between the lapse and the
1177: volume element (remember that now $\Theta=0$). It can be easily
1178: done in the case
1179: \begin{equation}\label{C_fdef}
1180:  f ~=~2\;\sigma~=~{\rm constant},
1181: \end{equation}
1182: namely
1183: \begin{equation}\label{C_dtAlpha}
1184:  \partial_t(\alpha \gamma^{-\sigma} ) ~=~ 0,
1185: \end{equation}
1186: so that the value of $\alpha$ can be defined in terms of $\gamma$ for every initial
1187: condition. The same thing can be done with $A_i$ and $d_i$, so that
1188: \begin{equation}\label{C_Adef}
1189:  A_i ~\equiv~\sigma\;{d_{ir}}^r~+~...
1190: \end{equation}
1191: and the set of dynamical fields is then further reduced to
1192: \begin{equation}\label{C_KSTuvector}
1193:  u ~ = ~ \{\gamma_{ij},~ K_{ij},~d_{kij} \}.
1194: \end{equation}
1195: The principal part of the evolution system is then given by (we
1196: suppress the tildes over the $K_{ij}$)
1197: 
1198: \begin{eqnarray}
1199: \label{BKSTdtgamma}
1200:  \partial_t \gamma_{ij} &=& \ldots
1201: \\
1202: \label{BKSTdtd}
1203:  \partial_t d_{kij} &+& \partial_r [
1204:   \; \alpha \{\,2\, \delta^r_k\, {K}_{ij} -\chi\;({{K}_{k}}^r-\delta^r_k\, \tr {K}) \gamma_{ij}
1205: \nonumber \\
1206:   &+&  \eta\;\gamma_{k(i}({{K}^r}_{j)}   - \delta^r_{j)} \tr {K})\, \}\;] = \ldots
1207: \\
1208: \label{BKSTdtK}
1209:  \partial_t K_{ij} &+& \partial_k~[\alpha~\lambda^k_{ij}] = \ldots
1210: \end{eqnarray}
1211: \begin{eqnarray}\label{BKSTdeflambda}
1212:    2\; {\lambda^k}_{ij} &=& {d^k}_{ij}
1213:    - \frac{n}{4}({d_{kr}}^r-{{d_r}^{rk}})\gamma_{ij}
1214: \nonumber \\
1215:    &-& \frac{1-\zeta}{2}(\delta^k_i\, {{d^r}_{rj}}+ \delta^k_j \, {{d^r}_{ri}})
1216:    + \frac{1+\zeta}{2}({d_{ij}}^k + {d_{ji}}^k )
1217: \nonumber \\
1218:    &+& \frac{1+2\sigma}{2}( \delta^k_i\, {d_{jr}}^r+ \delta^k_j\, {d_{ir}}^r)
1219: \end{eqnarray}
1220: 
1221: which corresponds precisely to (the principal part of) the original KST system \cite{KST01}.
1222: 
1223: {\em Acknowledgements: This work has been supported by the EU Programme
1224: 'Improving the Human Research Potential and the Socio-Economic
1225: Knowledge Base' (Research Training Network Contract HPRN-CT-2000-00137),
1226: by the Spanish Ministerio de Ciencia y Tecnologia through the research
1227: grant number BFM2001-0988 and by a grant from the Conselleria d'Innovacio
1228: i Energia of the Govern de les Illes Balears.}
1229: 
1230: \bibliographystyle{prsty}
1231: \begin{thebibliography}{99}
1232: 
1233: %\bibitem{Ch56} Y.~Choquet-Bruhat, J.~Rat.~Mee.~Analysis {\bf 5},
1234: %         951 (1956).
1235: \bibitem{Ligo} http://www.ligo.caltech.edu.
1236: 
1237: \bibitem{Virgo} http://www.virgo.infn.it.
1238: 
1239: \bibitem{Geo} http://www.geo600.uni-hannover.de.
1240: 
1241: \bibitem{Tama} http://tamago.mtk.nao.ac.jp.
1242: 
1243: \bibitem{GC} S.~Brandt et al, Phys.~Rev.~Lett. {\bf 85} 5496 (2000).
1244: 
1245: \bibitem{ADM62} R.~Arnowit, S.~Deser and C.~W.~Misner, {\em Gravitation:
1246:         an introduction to current research}, ed. L.~Witten,
1247:         Wiley, New York (1962).
1248: 
1249: \bibitem{Cent80} J.~Centrella, Phys.~Rev.~D {\bf 21}, 2776 (1980).
1250: 
1251: \bibitem{CR83} Y.~Choquet-Bruhat and T.~Ruggeri,
1252:          Comm.~Math.~Phys. {\bf 89}, 269 (1983).
1253: 
1254: \bibitem{BM92} C.~Bona and J.~Mass\'o,
1255:         Phys.~Rev.~Lett. {\bf 68} 1097 (1992)
1256: 
1257: \bibitem{BM95} C.~Bona, J.~Mass\'o, E.~Seidel
1258:         and J.~Stela, Phys.~Rev.~Lett. {\bf 75} 600 (1995).
1259: 
1260: \bibitem{SN95} M.~Shibata and T.~Nakamura,
1261:         Phys.~Rev.~D{\bf 52} 5428 (1995).
1262: 
1263: \bibitem{BS99} T.~W.~Baumgarte and S.~L.~Shapiro,
1264:         Phys.~Rev.~D{\bf 59} 024007 (1999).
1265: 
1266: \bibitem{FR94} S.~Frittelli and O.~A.~Reula,
1267:         Commun.~Math.~Phys. {\bf 166} 221 (1994).
1268: 
1269: \bibitem{FR96} S.~Frittelli and O.~A.~Reula,
1270:         Phys.~Rev.~Lett. {\bf 76} 4667 (1996).
1271: 
1272: \bibitem{AY99} A.~Anderson and J.~W.~York, Jr.,
1273:         Phys.~Rev.~Lett. {\bf 82} 4384 (1999).
1274: 
1275: \bibitem{Hern00} S.~D.~Hern, Ph.~D. Thesis, gr-qc/0004036.
1276: 
1277: \bibitem{KST01} L.~E.~Kidder, M.~A.~Scheel and
1278:          S.~A.~Teukolsky, Phys.~Rev. {\bf D64}, 064017 (2001).
1279: 
1280: \bibitem{ST02} O.~Sarbach and M.~Tiglio,  Phys.~Rev. {\bf D66},
1281:         064023 (2002).
1282: 
1283: \bibitem{AACY95} A.~Abrahams, A.~Anderson, Y.~Choquet-Bruhat and
1284:         J.~W.~York,  Phys.~Rev.~Lett. {\bf 75}, 3377 (1995).
1285: 
1286: \bibitem{Frie96} H.~Friedrich, Class.~Quantum.~Grav. {\bf 13}, 1451 (1996).
1287: 
1288: \bibitem{BFHR99} O.~Brodbeck, S.~Frittelli, P.~Hubner and
1289:         O.~Reula, J.~Math.~Phys. 40, 909-923 (1999).
1290: 
1291: \bibitem{SY} H.~Shinkai and G.~Yoneda ,``Progress in Astronomy and
1292:         Astrophysics" (Nova Science Publ), to be published. Also available
1293:         in gr-qc/0209111
1294: 
1295: \bibitem{BeA} C.~Bona et al, in preparation.
1296: 
1297: \bibitem{GGKM03} A.~P.~Gentle,N.~D.~George,A.~Kheyfets and W.~A.~Miller, gr-qc/0307007.
1298: 
1299: \bibitem{Z4} C.~Bona, T.~Ledvinka, C.~Palenzuela, M.~\v Z\'a\v cek,
1300:         Phys.~Rev.~D {\bf 67}, 104005 (2003).
1301: 
1302: \bibitem{KO01} H.~O.~Kreiss and O.~E.~Ortiz, Lect.~Notes Phys. 604 (2002) 359.
1303: 
1304: \bibitem{KL89} H.~O.~Kreiss and J.~Lorentz, ``Initial-Boundary Value Problems and
1305:         the Navier-Stokes Equations" in Pure and Applied Mathematics, Vol. 136
1306:   (Academic Press Inc., San Diego 1989).
1307: 
1308: \bibitem{Reu03} O.~Reula, Talk at the Hyperbolic Models for Astrophysics and
1309:        Cosmology Workshop (submitted to Phys.~Rev.~D)
1310: 
1311: \bibitem{Z3} C.~Bona, T.~Ledvinka and C.~Palenzuela,
1312:         Phys.~Rev.~D {\bf 66}, 084013 (2002).
1313: 
1314: \bibitem{Mexico} M.~Alcubierre et al, gr-qc/0305023.
1315: 
1316: \bibitem{DeDo21} T.~De Donder,{\em La Gravifique Einstenienne}
1317:        Gauthier-Villars, Paris (1921).
1318: 
1319: \bibitem{Fock59} Fock, V.A., {\em The theory of Space Time and Gravitation},
1320:        Pergamon, London (1959).
1321: 
1322: \bibitem{MoL} O.~A.~Liskovets, Differential equations I 1308-1323 (1965).
1323: 
1324: \bibitem{Gowdy71} R.~H.~Gowdy, Phys.~Rev.~D {\bf 27}, 826 (1971).
1325: 
1326: \bibitem{Berger} B.~Berger, submitted to Phys.~Rev.~D (2002),
1327:          gr-qc/0207035.
1328: 
1329: \bibitem{BM83} C.~Bona and J.~Mass\'o, Phys.~Rev. {\bf D38 } 2419 (1988).
1330: 
1331: \end{thebibliography}
1332: 
1333: \end{document}
1334: