gr-qc0201086/ccm
1: %=====================================================================
2: \section{Cauchy characteristic matching in cylindrical symmetry}
3: \label{ccm}
4: %
5: %
6: 
7: %=========================================================================
8: \subsection{The idea of Cauchy characteristic matching}
9: \label{CCM}
10: %
11: %
12: %
13: \begin{figure}[b]
14:   \centering
15:   \epsfig{file=CCMdemo.ps, height=300pt, width=350pt, angle=0}
16:   \caption{In this conformal diagram Cauchy characteristic
17:            matching is schematically illustrated.
18:            In the inner region matter is evolved with
19:            a ``3+1'' scheme whereas characteristic coordinates
20:            defined by the null geodesics are used in the
21:            outer vacuum region. The two formalisms are matched
22:            at the interface at finite radius.}
23:   \label{CCM_illu}
24: \end{figure}
25: %
26: Cauchy characteristic matching (CCM) is a method that simultaneously
27: makes use of the beneficial properties of the ``3+1'' and the characteristic
28: formalism. In section \ref{threep1} we have seen that in the ``3+1''
29: case spacetime is decomposed into 3-dimensional
30: space-like hypersurfaces threaded by a one parameter family of curves. The
31: dynamic variables are the
32: components $\hbox{\grvec g}_{ij}$ of the 3-metric of the hypersurfaces.
33: A complete set of initial data
34: consists of values for $\hbox{\grvec g}_{ij}$ and their time derivatives
35: on some initial hypersurface. The second order evolution equations then
36: determine the 4-metric of the spacetime up to gauge transformations.
37: This type of initial value problem is known as
38: a {\em Cauchy problem} and has been extensively used for the numerical
39: solution of Einstein's field equations.
40: It is however not suitable for the
41: analysis of gravitational radiation since it is not clear how to
42: incorporate null infinity into a finite numerical grid via conformal
43: compactification. Instead one uses approximating techniques to
44: extract information about the gravitational radiation at finite
45: radii and imposes {\em outgoing radiation
46: boundary conditions} in order to prevent incoming
47: gravitational waves. Unfortunately attempts to implement
48: these boundary conditions give rise
49: to spurious reflected numerical waves. Characteristic formalisms solve
50: this problem in an elegant way. Spacetime is decomposed into a
51: 2-parameter family of 2-dimensional space-like surfaces threaded by two
52: 1-parameter families of curves. At least one of these families
53: consists of null geodesics, the {\em characteristics} of the propagation
54: of radiation. The spacetime can be compactified by standard methods,
55: exact boundary conditions
56: can be applied at future or past null infinity and gravitational radiation
57: can be properly analysed. In regions of strong curvature, however, caustics
58: can form and the foliation along null geodesics breaks down.\\
59: A possible remedy for this problem
60: consists in using both a ``3+1'' and a characteristic formulation,
61: each in its preferred region.
62: Normally an astrophysical scenario is approximated as a finite inner region
63: containing all the matter (a neutron star, for example) and the outer
64: vacuum region with an observer located at future null infinity. In CCM
65: a ``3+1'' scheme is used for the evolution of the interior and
66: a characteristic formulation for the evolution of the exterior region.
67: At a finite radius an interface facilitates the transfer of information
68: between these two regions. The method is illustrated
69: in Fig.\,\ref{CCM_illu} where the dark shaded area represents the
70: astrophysical source. Gravitational waves emitted from this source
71: travel along null geodesics which are given by straight lines at an angle
72: of 45 degrees in this figure. In the outer region the null geodesics
73: are used to define the characteristic coordinate axis. \\
74: The feasibility of combining Cauchy algorithms with characteristic
75: methods in order to evolve the gravitational field was first studied
76: by \lcite{Bishop1992}. The first attempts at obtaining numerical evolutions
77: have been carried out in one spatial dimension. The work of
78: the Southampton CCM-group in cylindrical symmetry will be discussed in
79: detail in the next section. The Pittsburgh relativity
80: group studied CCM in spherical symmetry by evolving the
81: Einstein-Klein-Gordon system
82: (\shortciteNP{Gomez1996}). They have demonstrated second order
83: convergence and found no indications of back reflection or instabilities
84: at the interface. After the demonstration of the viability
85: of CCM in one dimension attention shifted towards higher-dimensional problems.
86: The Southampton relativity group focused their studies on the
87: axisymmetric case. After laying the theoretical foundations
88: (\citeNP{dInverno1996}, \citeNP{dInverno1997}) a great deal
89: of work has gone into the development of an
90: axisymmetric CCM code (see \citeNP{Pollney2000} for details). This
91: code has now been completed and is currently being evaluated and tested.
92: In contrast the Pittsburgh group has immediately turned their
93: attention towards the general 3-dimensional case.
94: \scite{Bishop1996} and \scite{Bishop1997} have probed the use
95: of Cauchy-characteristic matching in three dimensions by evolving
96: non-linear scalar waves in a flat space-time.
97: The application of these ideas to 3-dimensional problems
98: in general relativity has resulted in a module for the combination
99: of Cauchy and characteristic codes for the evolution of a binary black
100: hole (\shortciteNP{Bishop1998}). A more comprehensive overview of the
101: ongoing research using Cauchy-characteristic matching can be found in
102: \lcite{Winicour2001}.
103: 
104: 
105: %======================================================================
106: \subsection{The Southampton CCM-project}
107: %
108: %
109: The Southampton CCM-project is a long term project devoted to the
110: study of Cauchy-charac-teristic matching in scenarios of
111: decreasing symmetry assumptions
112: (\citeNP{dInverno2000}). The first step was to demonstrate
113: the viability of the approach. That was done by \citeN{Clarke1994} by
114: evolving the wave equation in flat spacetime. Attention then turned
115: towards gravitational waves in cylindrical symmetry. The theoretical 
116: foundations were laid by \shortciteN{Clarke1995} and the resulting
117: code of \shortciteN{Dubal1995}
118: showed good agreement with analytic solutions containing one gravitational
119: degree of freedom. Furthermore \shortciteANP{Dubal1995} demonstrated the
120: superior performance
121: of the CCM-method as compared with the use of artificial outer boundary
122: conditions in ``3+1'' schemes. \scite{dInverno2000b} presented
123: a generalisation of this
124: code to also include the rotational degree of freedom. They find,
125: however, that the convergence of the code drops to first
126: order level in later stages of the evolutions. In this work we will present
127: a new code that allows us
128: to include the rotational degree of freedom in terms of natural geometrical
129: variables with regular behaviour at null infinity. This reformulation
130: resulted in improved accuracy, long term stability and
131: ensures second order convergence over long evolution times.
132: We will demonstrate the improved quality by comparing the numerical results
133: with analytic solutions possessing both gravitational degrees of freedom. \\
134: The Southampton CCM-project has continued meanwhile with the
135: development of the axisymmetric code mentioned in the previous section.
136: %The latest stage in the Southampton CCM-project is the axisymmetric
137: %code mentioned in the . This code has been the subject
138: %of intensive work in recent years (see \citeNP{Pollney2000} for a
139: %detailed description) and is currently being evaluated and tested.
140: 
141: 
142: 
143: %======================================================================
144: \subsection{The original code} \label{CCM_org}
145: %
146: %
147: In this section we will describe the cylindrically symmetric Cauchy
148: characteristic matching code
149: developed by the Southampton Relativity Group (\shortciteNP{Clarke1995},
150: \shortciteNP{Dubal1995}). This code was used to reproduce the analytic
151: solution by \lcite{Weber1957}, which possesses one gravitational
152: degree of freedom,
153: with high accuracy and second order convergence. \scite{dInverno2000b}
154: presented an extension of this code based on the formulation
155: of \shortciteANP{Clarke1995} to also include the rotational degree of freedom.
156: Their difficulties in obtaining a long term stable second order convergent
157: code motivated the reinvestigation of the problem described in this thesis. \\
158: %The reformulation we will present below has resulted in a significant
159: %simplification of the
160: %relations at the interface. A new code has been written to implement
161: %these ideas and is compared with analytic solutions with one or two
162: %gravitational degrees of freedom. \\
163: In their derivation of the equations \shortciteANP{Clarke1995} find
164: it necessary to decompose spacetime according to the methods
165: of \lcite{Geroch1970} in order to eliminate irregularities of the
166: equations in the characteristic region.
167: The Geroch decomposition plays a crucial role in our reformulation
168: and will also be used in section \ref{cstr}
169: when we numerically simulate cosmic strings. 
170: Before we turn our attention to the cylindrically symmetric CCM code,
171: we will therefore describe the Geroch decomposition in more detail.
172: 
173: 
174: %========================================
175: \subsubsection{The Geroch decomposition}
176: \label{Gerochdec}
177: %
178: %
179: A problem generally faced in cylindrical symmetry is that the spacetime
180: is not asymptotically flat due to the infinite extension in the $z$-direction.
181: The decomposition of \lcite{Geroch1970} solves this problem by factoring out the
182: Killing direction and reformulating the 4-dimensional problem in terms
183: of two scalar fields on an asymptotically flat 3-dimensional spacetime.
184: Suppose, the spacetime admits a Killing 
185: field $\hbox{\grvec x}^{\mu}$ which in
186: the case of cylindrical symmetry simply is $\bbox{\partial}_z$.
187: Then we define the norm of the Killing vector
188: %
189: \begin{align}
190:   \nu &= \hbox{\grvec x}^{\alpha} \hbox{\grvec x}_{\alpha},
191:   \label{Geroch_norm}
192: \end{align}
193: %
194: and the Geroch twist
195: %
196: \begin{align}
197:   \hbox{\grvec t}_{\alpha} &= -\hbox{\grvec e}_{\alpha \beta \rho \sigma}
198:                    \hbox{\grvec x}^{\beta} 
199:                    \nabla^{\rho} \hbox{\grvec x}^{\sigma},
200:   \label{GEROCH_TWIST}
201: \end{align}
202: %
203: where $\hbox{\grvec e}_{\alpha \beta \rho \sigma}$ is the completely
204: antisymmetric Levi-Cevita tensor.
205: These fields are well defined on the 3-dimensional space $\mathcal{S}$
206: given by $z=\mathrm{const}$
207: with the resulting metric
208: %
209: \begin{align}
210:   \hbox{\vec h}_{\alpha \beta} &= \hbox{\vec g}_{\alpha \beta}
211:   - \frac{1}{\nu} \hbox{\grvec x}_{\alpha} \hbox{\grvec x}_{\beta}.
212:   \label{Geroch_3m}
213: \end{align}
214: %
215: If $D_{\mu}$ denotes the covariant derivative associated with the
216: 3-metric $\hbox{\vec h}$, one can show that
217: %
218: \begin{align}
219:   D_{[\rho}\hbox{\grvec t}_{\sigma ]} &= \hbox{\grvec e}_{\rho \sigma
220:       \alpha \beta} \hbox{\grvec x}^{\alpha}
221:       \hbox{\vec R}^{\beta}_{\,\,\, \lambda} \hbox{\grvec x}^{\lambda}.
222:       \label{Geroch_cond}
223: \end{align}
224: %
225: In vacuum the right hand side vanishes so that $\hbox{\grvec t}_{\sigma}$
226: is curl free and can be expressed in terms of a potential
227: %
228: \begin{align}
229:   \hbox{\grvec t}_{\sigma} &= D_{\sigma} \tau.
230:   \label{Geroch_pot}
231: \end{align}
232: %
233: It is a remarkable fact that the right hand side of Eq.\,(\ref{Geroch_cond})
234: will also vanish in some non-vacuum cases. In the discussion of cosmic strings
235: in section \ref{cstr} we will encounter such an example.\\
236: Geroch has then shown that the Einstein equations for the metric
237: $\hbox{\vec g}$ of the
238: 4-dimensional \mbox{spacetime} can be written in terms of the two scalar fields
239: $\tau$ and $\nu$ and the 3-metric $\hbox{\vec h}$
240: %
241: \begin{eqnarray}
242:  {\cal R}_{ab} &=& \frac{1}{2}\nu^{-2}[(D_a\tau)
243:                    (D_b\tau)-\hbox{\vec h}_{ab}(D_m\tau)
244:                    (D^m\tau)] + \frac{1}{2}\nu^{-1}
245:                    D_aD_b\nu - \frac{1}{4}\nu^{-2}(D_a\nu)
246:                    (D_b\nu)\nonumber\\
247:                 && +8\pi \hbox{\vec h}_a{}^\alpha \hbox{\vec h}_b{}^{\beta}
248:                    (\hbox{\vec T}_{\alpha\beta}-\frac{1}{2}
249:                    \hbox{\vec g}_{\alpha\beta}\hbox{\vec T}),
250:                    \label{Geroch3d} \\[10pt]
251:  D^2 \nu &=& \frac{1}{2}\nu^{-1}(D_m\nu) (D^m\nu) - \nu^{-1}(D_m\tau)
252:                       (D^m\tau) +16\pi (\hbox{\vec T}_{\alpha\beta}-
253:                       \frac{1}{2}\hbox{\vec g}_{\alpha\beta}\hbox{\vec T})
254:                       \hbox{\grvec x}^\alpha\hbox{\grvec x}^\beta, 
255:              \label{D91}\\[10pt]
256:  D^2 \tau &=& \frac{3}{2}\nu^{-1}(D_m\tau)
257:              (D^m\nu), \label{Geroch_tau}
258: \end{eqnarray}
259: %
260: where Latin indices run from 0 to 2 and $\mathcal{R}_{ab}$
261: is the Ricci tensor associated with the 3-metric $\hbox{\vec h}$.
262: Note that even in the
263: case of a vanishing energy-momentum tensor $\hbox{\vec T}$,
264: the scalar fields $\nu$ and $\tau$
265: present source terms in the field equations
266: (\ref{Geroch3d}) for the 3-metric $\hbox{\vec h}$. \\
267: In the vacuum case $\hbox{\vec T}_{\alpha \beta}=0$, \scite{Sjodin2000}
268: have shown how
269: it is possible to reformulate the Einstein-Hilbert
270: Lagrangian in terms of $\nu$, $\tau$ and the conformal 3-metric
271: $\tilde{\hbox{\vec h}}_{ab} = \nu \hbox{\vec h}_{ab}$.
272: This leads directly to the 3-dimensional energy-momentum tensor
273: %
274: \begin{equation}
275: \mathcal{T}_{ab} = \frac{1}{2}\nu^{-2}[{\tilde D}_a \tau{\tilde
276:                    D_b}\tau-\frac{1}{2}\tilde{\hbox{\vec h}}_{ab}
277:                    \tilde{ \hbox{\vec h}}^{cd}(\tilde
278:                    D_c\tau)(\tilde D_d\tau)
279:                    +{\tilde D}_a\nu{\tilde
280:                    D_b}\nu-\frac{1}{2}\tilde{ \hbox{\vec h}}_{ab}
281:                    \tilde{ \hbox{\vec h}}^{cd}(\tilde
282:                    D_c\nu)(\tilde D_d\nu)]
283:                    \label{Geroch_confT},
284: \end{equation}
285: %
286: where $\tilde{D}_a$ is the covariant derivative associated with the
287: conformal 3-metric $\tilde{\hbox{\vec h}}$.
288: Since the Weyl-curvature vanishes identically in three dimensions, the
289: curvature is completely determined by the Ricci tensor $\mathcal{R}_{ab}$,
290: i.e. the energy-momentum tensor $\mathcal{T}_{ab}$ which in turn is
291: determined by $\nu$ and $\tau$. Thus the gravitational
292: degrees of freedom of the original 4-dimensional spacetime are represented
293: by the scalar fields $\nu$ and $\tau$.
294: If matter is present in the
295: 4-dimensional spacetime, there are extra terms on the right hand side of
296: (\ref{Geroch_confT}).
297: 
298: 
299: %=======================================================================
300: \subsubsection{The equations of the original code}
301: \label{Dubal_equations}
302: %
303: %
304: We will now turn our attention to the original cylindrically symmetric
305: CCM code of the Southampton relativity group. An extensive description
306: of this code and the derivation of the equations can be found in
307: \scite{Clarke1995} and \scite{Dubal1995}. In order to illustrate the
308: effects of our reformulation, we will include here a rather
309: detailed description of their equations and choice of variables.
310: They start with the metric in Jordan, Ehlers, Kundt and
311: Kompaneets (JEKK) form (\shortciteNP{Jordan1960}, \citeNP{Kompaneets1958})
312: %
313: \begin{align}
314:   ds^2 &= e^{2(\gamma - \psi)}(-dt^2+dr^2) + r^2e^{-2\psi}d\phi^2
315:           +e^{2\psi} (\omega d\phi + dz)^2,
316:   \label{JEKK}
317: \end{align}
318: %
319: which describes a general cylindrically symmetric vacuum spacetime. The
320: metric functions $\psi$, $\omega$ and $\gamma$ are functions of $(r,t)$
321: only. In terms of the gauge freedom discussed in section \ref{ADM_GAUGE}
322: this choice implies a vanishing shift vector and the lapse is determined
323: by the requirement that $\hbox{\vec g}_{tt}=-\hbox{\vec g}_{rr}$. As a
324: consequence the null geodesics are given by the simple relations
325: $t\pm r = \mathrm{const}$.
326: In the outer characteristic region, the line element is rewritten by
327: transforming to the coordinates
328: %
329: \begin{align}
330:   u &= t-r, \\[10pt]
331:   y &= \frac{1}{\sqrt{r}},
332:   \label{y}
333: \end{align}
334: %
335: and the regions are matched at $r=1=y$. \shortciteANP{Clarke1995} 
336: find, however, that the
337: compactified field equations cannot be made regular in this way.
338: Therefore they factor out the $z$-direction in the outer region according to
339: the Geroch decomposition described above. This leads to
340: a reformulation of the problem in terms of the variables
341: %
342: \begin{align}
343:   m &= \frac{\nu - 1}{y}, \label{DUBAL_M} \\[10pt]
344:   w &= \frac{\tau}{y},
345: \end{align}
346: %
347: where $\tau$ is the Geroch potential and $\nu$ the
348: norm of the $z$-Killing vector. These are related to the metric functions
349: $\psi$ and $\omega$ by Eqs.\,(\ref{Geroch_norm}) and (\ref{GEROCH_TWIST})
350: which in this particular case become
351: %
352: \begin{align}
353:   \nu &= e^{2\psi}, \label{DUBAL_PSI} \\[10pt]
354:   \tau_{,y} &= y^2 e^{4\psi} \omega_{,u}.
355: \end{align}
356: %
357: With this choice of variables one obtains two
358: evolution equations for $\psi$ and $\omega$ in the interior
359: Cauchy region and a constraint equation for $\gamma$. \shortciteANP{Dubal1995}
360: write this set of equations as a first order system
361: %
362: \begin{align}
363:   \psi_{,t} &= \frac{1}{r} \tilde{L}, \label{DUBAL_PSIT} \\[10pt]
364:   \omega_{,t} &= -2 e^{-4\psi} L^{\phi}_z, \label{DUBAL_OMEGAT} \\[10pt]
365:   \tilde{L}_{,t} &= \frac{1}{r}[ r^2\psi_{,rr} + r\psi_{,r} -
366:      \frac{1}{2} e^{4\psi} \omega^2_{,r} + 2 e^{-4\psi}(L^{\phi}_z)^2],
367:      \label{DUBAL_LT} \\[10pt]
368:   L^{\phi}_{z,t} &= \frac{1}{r}e^{4\psi} (\frac{1}{2} \omega_{,r} -
369:      \frac{1}{2} r \omega_{,rr} - 2r \psi_{,r} \omega_{,r}),
370:      \label{DUBAL_LPZT} \\[10pt]
371:   \chi_{,r} &= \frac{1}{4r} e^{4\psi} \omega^2_{,r} - \psi_{,r} +
372:      r\psi^2_{,r} + \frac{1}{r}[\tilde{L}^2 + e^{-4\psi}(L^{\phi}_z)^2],
373:     \label{DUBAL_CHIR}
374: \end{align}
375: %
376: where $\chi=\gamma - \psi$. The corresponding set of equations in the
377: characteristic region is given by two evolution equations for $m$ and $w$
378: and a hypersurface equation for $\gamma$ which is again written as a first
379: order system
380: %
381: \begin{align}
382:   m_{,u} = \,\,&\nu M,     \label{DUBAL_MU} \\[10pt]
383:   w_{,u} = \,\,&\nu W,     \label{DUBAL_WU} \\[10pt]
384:   \begin{split}
385:   M_{,y} = &-\frac{1}{\nu}(yw)_,{y}W + \frac{1}{4\nu}\biggl[
386:             -y(m+y^2m_{,yy} + 3ym_{,y})\\[10pt]
387:            & +\frac{1}{\nu}y^2(m^2
388:             +2ymm_{,y} - w^2 - 2yww_{,y} + y^2m^2_{,y} -
389:             y^2w^2_{,y})\biggr], \label{DUBAL_MY}
390:   \end{split} \\[10pt]
391:   \begin{split}
392:   W_{,y} = \,\,&\frac{1}{\nu}(yw)_{,y} M + \frac{1}{4\nu} \biggl[
393:             -y(w+y^2w_{,yy}+3yw_{,y})\\[10pt]
394:             & +\frac{1}{\nu}2y^2
395:             (mw+ymw_{,y}+ywm_{,y}+y^2m_{,y}+y^2m_{,y}w_{,y}) \biggr],
396:             \label{DUBAL_WY}
397:   \end{split} \\[10pt]
398:   \gamma_{,y} = & -\frac{1}{8\nu^2}y[m^2+w^2+2y(mm_{,y}+ww_{,y})
399:             +y^2(m^2_{,y} + w^2_{,y})]. \label{DUBAL_GAMMAY}
400: \end{align}
401: %
402: The transformation between the two pairs
403: of variables $(\psi, \omega)$ and $(m,w)$ and their derivatives is
404: implemented at the interface at $r=1=y$ according to the relations
405: %
406: \begin{eqnarray}
407:   \tilde{L} &=& \frac{M}{2y}, \label{DUBAL_TILDEL} \\[10pt]
408:   \omega_{,r} &=& \frac{W}{y\nu}, \label{DUBAL_OMEGAR}\\[10pt]
409:   m &=& \frac{1}{y} (e^{2\psi}-1),\label{Dubal_m}  \\[10pt]
410:   (yw)_{,y} &=& -\frac{2}{r} L_z^{\phi}. \label{DUBAL_YWY}
411: \end{eqnarray}
412: %
413: The problematic relations are (\ref{DUBAL_OMEGAR}) and
414: (\ref{DUBAL_YWY}) which
415: involve the spatial derivative of $\omega$ and $w$.
416: The presence of spatial derivatives in combination with the
417: interpolation techniques applied at the interface make the implementation
418: of these relations a rather subtle issue. \\
419: The code of \shortciteANP{Dubal1995} formed the starting point for our
420: investigation of the problem. This code has been
421: well checked in the non-rotating case but did not include the implementation
422: of Eqs.\,(\ref{DUBAL_OMEGAR}) and (\ref{DUBAL_YWY}) for
423: the rotational variables
424: $\omega$ and $w$ at the interface. In this work we therefore started with
425: the addition of these missing modules to the original code. In order to
426: describe our implementation it is necessary to first discuss the numerical
427: techniques, in particular those underlying the transmission of
428: information from the Cauchy to the characteristic region and vice versa.
429: 
430: 
431: %========================================================================
432: \subsubsection{The numerical implementation}
433: \label{start_int}
434: %
435: %
436: We will now discuss the numerical implementation of
437: Eqs.\,(\ref{DUBAL_PSIT})-(\ref{DUBAL_YWY}).
438: %has been described in detail in \shortciteNP{Dubal1995}. Because of the
439: %importance of the numerical methods in particular at the interface, we
440: %summarise of the essential steps.
441: The numerical grid used for the evolution consists of an inner
442: Cauchy region which covers the range $0 \leq r \leq 1$ and the 
443: outer characteristic region extending from $r=1$ to infinity which 
444: corresponds to the range $1 \geq y \geq 0$.
445: The evolution equations in these regions
446: are discretized in a straightforward way using the leapfrog scheme
447: described in section \ref{FDE_LEAP} while second order centered finite
448: differencing is used for the constraints. 
449: If we assume that all functions are known on the time slices $n$, $n-1$ and
450: $n-2$, a full evolution cycle consists of the following steps.
451: %
452: \begin{list}{\rm{(\arabic{count})}}{\usecounter{count}
453:              \labelwidth1cm \leftmargin1.5cm \labelsep0.4cm \rightmargin1cm
454:              \parsep0.5ex plus0.2ex minus0.1ex \itemsep0ex plus0.2ex}
455:   \item Evolution of $\psi$, $\omega$, $\tilde{L}$ and $L_z^{\phi}$
456:         at the interior grid points of the Cauchy region according to
457:         Eqs.\,(\ref{DUBAL_PSIT})-(\ref{DUBAL_LPZT}).
458:   \item Update of these variables at the origin according to the inner
459:         boundary conditions $\psi_{,r}=\omega_{,r}=\tilde{L}_{,r}
460:         =L^{\phi}_{z,r}=0$.
461:   \item Evolution of $\psi$ and $\omega$ at the outer boundary of the Cauchy
462:         grid ($r=1$) according to Eqs.\,(\ref{DUBAL_PSIT}),
463:         (\ref{DUBAL_OMEGAT}).
464:   \item Extraction of $\psi$ and $\omega$ from the interface at $1+dr$ on
465:         the Cauchy grid on time slice $n$.
466:   \item Evolution of $\tilde{L}$, $L_z^{\phi}$ at the outer boundary
467:         of the Cauchy grid ($r=1$) according to
468:         Eqs.\,(\ref{DUBAL_LT}), (\ref{DUBAL_LPZT}).
469:   \item Calculation of $\chi$ on the Cauchy grid via quadrature according to
470:         Eq.\,(\ref{DUBAL_CHIR}).
471:   \item Evolution of $m$ and $w$ in the characteristic region according to
472:         Eqs.\,(\ref{DUBAL_MU}), (\ref{DUBAL_WU}).
473:   \item Extraction of $m$ and $w$ from the interface at $1+dy$ on the
474:         characteristic grid on time slice $n+1$.
475:   \item Calculation of $M$, $W$ and $\gamma$ on the characteristic grid
476:         via quadrature according to Eqs.\,(\ref{DUBAL_MY})-(\ref{DUBAL_GAMMAY}).
477: \end{list}
478: %
479: The crucial steps which provide the flow of information through the
480: interface are (4) and (8). These steps together with the start up procedure
481: required to get the leap-frog scheme running will now be discussed in more
482: detail. We start with the interface. \\
483: We first note that the interface is fixed at the radial position $r=1=y$.
484: Since we always have the freedom to rescale the radial coordinate $r$ by
485: a constant factor, this implies no loss of generality. From a numerical point
486: of view the need of an interface arises from the calculation of spatial
487: derivatives at $r=1$ on the Cauchy grid and $y=1$ on the characteristic grid.
488: The centred finite differencing used for the leapfrog scheme as illustrated
489: in Eq.\,(\ref{FDE_leapfrog}) requires knowledge of the Cauchy variables
490: at $r=1+dr$ and the characteristic variables at $y=1+dy$ for this purpose.
491: In order to obtain these values, they need to be calculated with interpolation
492: techniques using Eqs.\,(\ref{DUBAL_TILDEL})-(\ref{DUBAL_YWY}). We will
493: describe this process in the case of the direction ``char$\rightarrow$Cauchy''
494: corresponding to step (4). The reverse direction in step (8) works in
495: complete analogy. The situation is graphically illustrated in
496: Fig.\,\ref{Interface}. 
497: %
498: \begin{figure}
499:   \centering
500:   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
501: \begin{picture}(0,0)%
502: \epsfig{file=Interface3.pstex}%
503: \end{picture}%
504: \setlength{\unitlength}{3947sp}%
505: %
506: \begingroup\makeatletter\ifx\SetFigFont\undefined%
507: \gdef\SetFigFont#1#2#3#4#5{%
508:   \reset@font\fontsize{#1}{#2pt}%
509:   \fontfamily{#3}\fontseries{#4}\fontshape{#5}%
510:   \selectfont}%
511: \fi\endgroup%
512: \begin{picture}(6087,4737)(301,-4036)
513: \put(1201,-2761){\makebox(0,0)[lb]{\smash{\SetFigFont{12}{14.4}{\familydefault}{\mddefault}{\updefault}$dt$}}}
514: \put(6301,-961){\makebox(0,0)[lb]{\smash{\SetFigFont{12}{14.4}{\rmdefault}{\mddefault}{\updefault}$-y$}}}
515: \put(2176,-2161){\makebox(0,0)[lb]{\smash{\SetFigFont{12}{14.4}{\rmdefault}{\mddefault}{\updefault}$r_{K-1}$}}}
516: \put(3076,-2161){\makebox(0,0)[lb]{\smash{\SetFigFont{12}{14.4}{\rmdefault}{\mddefault}{\updefault}$r_K$}}}
517: \put(2851,-286){\makebox(0,0)[lb]{\smash{\SetFigFont{12}{14.4}{\rmdefault}{\mddefault}{\updefault}$r=1=y$}}}
518: \put(3376,-736){\makebox(0,0)[lb]{\smash{\SetFigFont{12}{14.4}{\rmdefault}{\mddefault}{\updefault}$t,u$}}}
519: \put(5176,-886){\makebox(0,0)[lb]{\smash{\SetFigFont{12}{14.4}{\rmdefault}{\mddefault}{\updefault}$du$}}}
520: \put(2701,-3736){\makebox(0,0)[lb]{\smash{\SetFigFont{12}{14.4}{\rmdefault}{\mddefault}{\updefault}$dr$}}}
521: \put(3676,-3661){\makebox(0,0)[lb]{\smash{\SetFigFont{12}{14.4}{\rmdefault}{\mddefault}{\updefault}$dy$}}}
522: \put(901,-4036){\makebox(0,0)[lb]{\smash{\SetFigFont{12}{14.4}{\rmdefault}{\mddefault}{\updefault}$-r$}}}
523: \put(4276,-2236){\makebox(0,0)[lb]{\smash{\SetFigFont{12}{14.4}{\rmdefault}{\mddefault}{\updefault}$r_{K+1}$}}}
524: \put(301,-3886){\makebox(0,0)[lb]{\smash{\SetFigFont{12}{14.4}{\rmdefault}{\mddefault}{\updefault}$n-2$}}}
525: \put(301,-3136){\makebox(0,0)[lb]{\smash{\SetFigFont{12}{14.4}{\rmdefault}{\mddefault}{\updefault}$n-1$}}}
526: \put(301,-2386){\makebox(0,0)[lb]{\smash{\SetFigFont{12}{14.4}{\rmdefault}{\mddefault}{\updefault}$n$}}}
527: \end{picture}
528:   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
529:   \caption{The interface in the direction from the characteristic to the
530:            Cauchy region. See the text for details.}
531:   \label{Interface}
532: \end{figure}
533: %
534: The derivatives of a function $f$ at $r=1$ can be calculated
535: to second order accuracy by centred finite differencing
536: %
537: \begin{align}
538:   f_{,r}|_K &= \frac{f_{K+1} - f_{K-1}}{2 dr}, \label{Interface_fr} \\[10pt]
539:   f_{,rr}|_K &= \frac{f_{K+1} -2f_K + f_{K-1}}{dr^2}, \label{Interface_frr}
540: \end{align}
541: %
542: if $f_{K+1}$ is obtained from interpolation to fourth order accuracy
543: in the characteristic region. For this purpose $\psi$, $\omega_{,r}$,
544: $\tilde{L}$ and $L^{\phi}_z$
545: are calculated in terms of the characteristic variables according to
546: Eqs.\,(\ref{DUBAL_TILDEL})-(\ref{DUBAL_YWY}) at the 12 points
547: of the characteristic 
548: region (including 3 points at the interface) indicated by filled
549: circles in Fig.\,\ref{Interface}. These values can then
550: be used to obtain the function
551: values $\psi_{K+1}$ and $\omega_{K+1}$ at location $r_{K+1}$ with
552: the required accuracy. \\
553: An alternative to this method consists in using the
554: same interpolation technique to calculate
555: the $r$-derivatives $\psi_{,r}$ and
556: $\omega_{,r}$ at grid point $K+1$ instead of the function values $\psi$
557: and $\omega$. We can then 
558: calculate the $r$-derivatives at the interface from
559: %
560: \begin{align}
561:   f_{,r}|_K &= \frac{f_{,r}|_{K+1} + f_{,r}|_{K-1}}{2}, 
562:   \label{Interface_fr2} \\[10pt]
563:   f_{,rr}|_K &= \frac{f_{,r}|_{K+1} - f_{,r}|_{K-1}}{2 dr}.
564:   \label{Interface_frr2}
565: \end{align}
566: %
567: Even though this alternative looks natural for the transformation
568: between $\omega$ and $w$ because these variables are related via
569: their derivatives according to Eq.\,(\ref{DUBAL_OMEGAR}),
570: it does not lead to any improvement of the performance of the code. \\
571: The second point we need to discuss is the so-called start-up problem.
572: It is an intrinsic difficulty of 3-level schemes such as the leap-frog
573: algorithm that the specification of initial data on one time slice
574: will not be sufficient to start the numerical engine. Instead
575: different techniques need to be used to obtain data on auxiliary
576: time slices. Due to the requirements of the fourth-order interpolation
577: at the interface we need information on two additional slices. The data
578: on these auxiliary slices are calculated in three steps.
579: %
580: \begin{list}{\rm{(\arabic{count})}}{\usecounter{count}
581:              \labelwidth1cm \leftmargin1.5cm \labelsep0.4cm \rightmargin1cm
582:              \parsep0.5ex plus0.2ex minus0.1ex \itemsep0ex plus0.2ex}
583:   \item The first order Euler scheme (see for example \shortciteNP{Press1989})
584:         is used to calculate data at $t=t_0-dt/2$.
585:   \item This auxiliary time slice is then used to determine the variables
586:         at $t=t_0-dt$ according to the leapfrog scheme.
587:   \item In another leapfrog step, this time using the full time step $dt$,
588:         data is calculated at $t_0-2dt$.
589: \end{list}
590: %
591: An alternative treatment at the interface is required for this
592: start-up procedure, because
593: the necessary three time-slices are not available at this stage. For
594: this purpose the
595: Cauchy grid is extended into the characteristic region by 10 grid
596: points. The derivatives of the Cauchy variables
597: can thus be calculated at $r=1$ using
598: centred finite differencing and the derivatives of the characteristic
599: variables follow from chain-rule. The treatment of the outer boundary of the
600: Cauchy grid is irrelevant for the numerical evolution, since the spurious
601: signal cannot travel across the additional 10 grid points during the
602: three evolution steps at the start-up procedure and these points
603: are not used in the remaining evolution.  
604: 
605: 
606: %========================================================================
607: \subsubsection{Including the rotational degree of freedom $\omega$}
608: %
609: %
610: In our first attempt to include the rotational degree of freedom we have
611: made use of the set of variables of section \ref{Dubal_equations}, 
612: namely $\psi$, $\omega$ and $\chi$ in the inner 
613: and $m$, $w$ and $\gamma$ in the outer region. For this purpose
614: we have extended the interface of the original code
615: to also include the transformations
616: between $\omega$ and $w$ as described in the previous section.
617: In order to test the code we use the analytic solution from
618: \lcite{Xanthopoulos1986} which we will discuss in more detail
619: in section \ref{as}. In Eqs.\,(\ref{as_Q})-(\ref{as_ygamma}) we give
620: analytic expressions for this solution in terms of the
621: Killing vector $\nu$, the Geroch potential $\tau$
622: and the metric function $\gamma$. The corresponding results for the
623: variables $\psi$, $m$ and $w$ are obtained straightforwardly from their
624: definitions (\ref{DUBAL_M})-(\ref{DUBAL_PSI}). The transformation
625: into values for the function $\omega$ is more complicated. The result
626: is given by \scite{Sjodin2000}
627: %
628: \begin{align}
629:   \omega(t,r) &= \sqrt{a^2+1}(X+Q-2)\frac{Z-Y}{2aZ},
630: \end{align}
631: %
632: where the auxiliary functions $Q$, $X$, $Y$ and $Z$ are defined
633: in Eqs.\,(\ref{as_Q})-(\ref{as_Z}).
634: We have not been able, however, to obtain a long term stable evolution
635: in this formulation of the problem. 
636: For 300 grid points in each region
637: instability set in after less than 1000 time steps and from the
638: pattern of the noise it is clear that the problems originate at the
639: interface. In our attempts to overcome the instability we have
640: varied the obvious parameters such as the Courant factor and
641: the number of grid points over a large range, but no improvement
642: has been achieved. We have also used the alternative implementation
643: of the interface according to Eqs.\,(\ref{Interface_fr2}),
644: (\ref{Interface_frr2}). Even though this alternative looks quite natural
645: at least for the transformation between $\omega$ and $w$ which are
646: related via their derivatives according to (\ref{DUBAL_OMEGAR}), we
647: did not achieve a significantly better performance with this method.
648: Finally we have
649: changed the start time of the numerical evolution and, thus, the
650: initial data. The obvious choice $t=0$ is not possible because some
651: derivatives of Xanthopoulos' solution are discontinuous at $t=0$,
652: but any positive value large enough to ensure that the start-up procedure does
653: not extend to negative times can be chosen. Again the code
654: became unstable after less than 1000 time steps. We have therefore
655: decided to restart the investigation of this problem by looking
656: for alternative sets of variables.
657: 
658: 
659: %========================================================================
660: \subsection{A reformulation of the problem}
661: %
662: %
663: A striking peculiarity of the formulation described above is the
664: drastically different treatment of the Cauchy and the characteristic region.
665: In view of the numerical subtleties associated with the interface one may
666: question the wisdom of factoring out the $z$-direction in one region
667: and work in the framework of the 4-dimensional spacetime in the other.
668: It rather seems natural to look for as homogeneous a description of the whole
669: spacetime as possible. In this context it is worth noting that the
670: restriction of the Geroch decomposition to the characteristic region
671: was a voluntary choice and not enforced at any stage of the derivation
672: of the equations. We have therefore decided to factor out the $z$-direction
673: in the Cauchy region as well and thus Geroch decomposed the whole spacetime.
674: This enables us to use the same set of fundamental variables throughout
675: spacetime and thus obtain almost trivial interface relations. A closer
676: investigation of the equations suggests that aside from the metric function
677: $\gamma$ the geometric variables $\nu$ and $\tau$ are
678: the natural variables to describe the cylindrically symmetric spacetime.
679: With this choice the equations in the Cauchy region can be written as
680: %
681: \begin{align}
682:   \nu_{,tt} &= \frac{1}{\nu}(\nu_{,t}^2 -\nu_{,r}^2+\tau_{,r}^2-\tau_{,t}^2)
683:       +\nu_{,rr} +\frac{\nu_{,r}}{r}, \label{CCM_NUTT} \\[10pt]
684:   \tau_{,tt} &= \frac{2}{\nu}(\tau_{,t} \nu_{,t} - \tau_{,r} \nu_{,r})
685:       + \tau_{,rr}+\frac{1}{r}\tau_{,r}, \label{CCM_TAUTT} \\[10pt]
686:   \gamma_{,r} &= \frac{r}{4\nu^2}(\nu_{,r}^2 +\nu_{,t}^2+\tau_{,r}^2
687:       +\tau_{,t}^2). \label{CCMgammar}
688: \end{align}
689: %
690: In practice we use $\nu_{,t}$ and $\tau_{,t}$ as
691: auxiliary variables in order to write Eqs.\,(\ref{CCM_NUTT}),
692: (\ref{CCM_TAUTT}) as a first order system.
693: If we transform to the new set of variables 
694: the equations in the characteristic region become
695: %
696: \begin{align}
697:   \nu_{,u} &= y\nu M, \\[10pt]
698:   \tau_{,u} &= y\nu W, \\[10pt]
699:   M_{,y} &= -\frac{y}{4\nu} \left[y\nu_{,yy} + \nu_{,y} + \frac{y}{\nu}
700:             (\tau_{,y}^2 - \nu_{,y}^2)\right]-W\frac{\tau_{,y}}{\nu}, \\[10pt]
701:   W_{,y} &= -\frac{y}{4\nu} \left(y\tau_{,yy} + \tau_{,y}-2\frac{y}{\nu}\nu_{,y}
702:             \right) + M\frac{\tau_{,y}}{\nu}, \\[10pt]
703:   \gamma_{,y} &= -\frac{y}{8\nu^2}(\tau_{,y}^2 +\nu_{,y}^2). \label{CCMgammay}
704: \end{align}
705: %
706: Finally the non-trivial relations at the interface are now given by
707: %
708: \begin{align}
709:   \nu_{,t} &= y\nu M, \\[10pt]
710:   \tau_{,t} &= y\nu W.
711: \end{align}
712: %
713: We have developed a code using the numerical techniques of section
714: \ref{start_int} based on these evolution equations and interface 
715: relations.
716: % In the next section we will discuss the performance of this
717: % new scheme.
718: 
719: 
720: %======================================================================
721: \subsection{Testing the code}
722: %
723: %
724: In order to test the performance of the new code, we will check it against 
725: analytic solutions with one and two gravitational degrees of freedom.
726: Furthermore we will demonstrate its internal consistency with a
727: time dependent convergence analysis. \\
728: We have already mentioned the vacuum solution by \lcite{Weber1957} that
729: was successfully used by \shortciteANP{Dubal1995} 
730: to test their CCM code. A solution with
731: both gravitational degrees of freedom was derived by \lcite{Xanthopoulos1986}.
732: Both these solutions can be rewritten in terms of our
733: variables $\nu$, $\tau$ and $\gamma$ and thus compared with the numerical
734: results.
735: 
736: 
737: %====================================
738: \subsubsection{The Weber-Wheeler wave}
739: \label{CCM_ww}
740: %
741: %
742: The analytic solution by \citeANP{Weber1957} describes a gravitational
743: pulse of the ``$+$'' polarization mode that moves in from past
744: null infinity, implodes on the axis and emanates away to future
745: null infinity. The analytic expressions in terms of $\nu$
746: and $\gamma$
747: have been derived in \scite{Sjodin2000}. In the Cauchy region it
748: is convenient to introduce the auxiliary quantities
749: %
750: \begin{align}
751:   X &= a^2 + r^2 - t^2, \label{ww_X} \\[10pt]
752:   Y &= X^2+4a^2t^2,
753: \end{align}
754: %
755: and the Weber-Wheeler wave can be written as
756: %
757: \begin{align}
758:   \nu &= \exp \left[ 2b \sqrt{\frac{2(X+\sqrt{Y})} {Y}}
759:             \right], \\[10pt]
760:   \gamma &= \frac{b^2}{2a^2} \left[ 1 - 2a^2r^2 \frac{X^2-4a^2t^2}
761:             {Y^2} - \frac{a^2+t^2-r^2}{\sqrt{Y}} \right],
762:   \label{ww_gamma}
763: \end{align}
764: %
765: where $a$ and $b$ are constants representing the width and amplitude 
766: of the pulse. The corresponding result in terms of the characteristic 
767: coordinates $u$, $y$ is
768: %
769: \begin{align}
770:   \tilde{X} &= a^2 y^2 - u^2 y^2 - 2u, \label{ww_yX} \\[10pt]
771:   \tilde{Y} &= \tilde{X}^2 + 4a^2(uy^2+1), \\[10pt]
772:   \nu &= \exp \left[ 2by \sqrt{\frac{2(\tilde{X}
773:               +\sqrt{\tilde{Y}})} {\tilde{Y}}} \right], \\[10pt]
774:   \gamma &= \frac{b^2}{2a^2} \left[ 1 - 2a^2 \frac{\tilde{X}^2-4a^2(uy^2+1)^2}
775:             {\tilde{Y}^2} - \frac{a^2y^2+u^2y^2+2u}{\sqrt{\tilde{Y}}} \right].
776:   \label{ww_ygamma}
777: \end{align}
778: %
779: The initial values for $\nu$ and its time
780: derivative are prescribed according
781: to these equations whereas $\gamma$ on the initial slice is calculated 
782: via quadrature from the constraint equations (\ref{CCMgammar}) and
783: (\ref{CCMgammay}). In order
784: to plot the solution for $0\le r < \infty$ we introduce the radial variable
785: %
786: \begin{align}
787:   w &= \left\{ \parbox{4cm}
788:                {
789:                $r \hspace{1.21cm} {\rm for}\,\, 0\le r \le 1 \\[5pt]
790:                3-\frac{2}{\sqrt{r}} \hspace{0.27cm} {\rm for}\,\, r > 1. $
791:                } \right. \label{w}
792: \end{align}
793: %
794: In Fig.\,\ref{CCMng} we show the numerical results for $\nu$
795: and $\gamma$
796: and their deviation from the analytic values obtained for $a=2$ and $b=0.5$
797: using 1200 grid points in each region and a Courant factor of 0.45. 
798: As in the case of the original code from \shortciteANP{Dubal1995} we find that
799: a Courant factor $<0.5$ is required for a stable evolution.
800: The plots show the incoming pulse in $\nu$ which is reflected at
801: the origin and then moves outwards to null infinity. The relatively
802: large number of grid points is required to achieve a high accuracy 
803: at early times in modelling the steep gradients of the incoming pulse.
804: If the calculation starts at a later time or a smaller parameter $a$ for 
805: the width of the pulse is used, the same accuracy is obtained with
806: significantly fewer grid points. 
807: %
808: \begin{figure}[t]
809:   \centering
810:   \epsfig{file=ww_ccm_nu.ps, height=400pt, width=175pt, angle=-90}
811:   \epsfig{file=ww_ccm_gam.ps, height=400pt, width=175pt, angle=-90}
812:   \caption{The numerical solutions for $\nu$ and $\gamma$ of the
813:            Weber-Wheeler wave for $a=0.5$, $b=2$ obtained with
814:            1200 grid points in each region (left panels). In the right
815:            panels the corresponding deviation from the analytic result
816:            is amplified by $10^5$ and $10^6$, respectively. For presentation
817:            purposes $\nu$ and $\gamma$ are viewed from
818:            different angles.}
819:   \label{CCMng}
820: \end{figure}
821: %
822: We also see that longer runs do not reveal any 
823: new features as the metric variables approach their Minkowskian values
824: after $t\approx 5$. This solution, however,
825: does not provide a test for the rotational degree of freedom. For that purpose
826: we need an analytic solution with both gravitational degrees of freedom.
827: 
828: 
829: %=======================================================================
830: \subsubsection{Xanthopoulos' rotating solution}
831: \label{as}
832: %
833: %
834: The next solution we consider is one due to \scite{Xanthopoulos1986}
835: which has a conical singularity on the $z$-axis and therefore
836: describes a rotating vacuum solution with a cosmic string type
837: singularity. The solution has been rewritten in terms of our variables
838: by \scite{Sjodin2000}. Again it is convenient to introduce auxiliary
839: quantities
840: %
841: \begin{align}
842:     Q&=r^2-t^2+1, \label{as_Q} \\[10pt]
843:     X&=\sqrt{Q^2+4t^2}, \label{as_X} \\[10pt]
844:     Y&=\frac{1}{2}[(2a^2+1)X+Q]+1-a\sqrt{2(X-Q)},\\[10pt]
845:     Z&=\frac{1}{2}[(2a^2+1)X+Q]-1, \label{as_Z}
846: \end{align}
847: %
848: where $a$ is a free parameter which can take on any non-zero value.
849: The solution derived by Xanthopoulos then becomes
850: %
851: \begin{align}
852:     \nu(t,\rho)&=\frac{Z}{Y}, \\[10pt]
853:     \tau(t,\rho)&=-\frac{\sqrt{2(a^2+1)}\sqrt{X+Q}}{Y}, \\[10pt]
854:     \gamma(t,\rho)&=\frac{1}{2}\ln \frac{Z}{a^2X}.
855: \end{align}
856: %
857: In the outer region where we use the coordinates $(u,y)$ the result is
858: %
859: \begin{align}
860:     \tilde{Q} &= y^2-u^2y^2-2u, \\[10pt]
861:     \tilde{X} &= \sqrt{\tilde{Q}^2+4(uy^2+1)^2}, \\[10pt]
862:     \tilde{Y} &= \frac{1}{2}[(2a^2+1)\tilde{X}+\tilde{Q}]+y^2
863:                  -ay\sqrt{2(\tilde{X}-\tilde{Q})}, \\[10pt]
864:     \tilde{Z} &= \frac{1}{2}[(2a^2+1)\tilde{X}+\tilde{Q}]-y^2, \\[10pt]
865:     \nu(u,y)  &= \frac{\tilde{Z}}{\tilde{Y}}, \\[10pt]
866:     \tau(u,y) &= -\frac{\sqrt{2(a^2+1)}\sqrt{\tilde{X}
867:                  +\tilde{Q}}}{\tilde{Y}}, \\[10pt]
868:     \gamma(u,y) &= \frac{1}{2}\ln \frac{\tilde{Z}}{a^2\tilde{X}}.
869:     \label{as_ygamma}
870: \end{align}
871: %
872: In Fig.\,\ref{CCMntg} we show the numerical results and the deviation from the
873: %
874: \begin{figure}[t]
875:   \centering
876:   \epsfig{file=as_ccm_nu.ps, height=400pt, width=175pt, angle=-90}
877:   \epsfig{file=as_ccm_tau.ps, height=400pt, width=175pt, angle=-90}
878:   \epsfig{file=as_ccm_gam.ps, height=400pt, width=175pt, angle=-90}
879:   \caption{The numerical solutions for $\nu$,
880:            $\tau$ and $\gamma$ of
881:            Xanthopoulos' spacetime for $a=1$ obtained with
882:            300 grid points in each region (left panels). In the right
883:            panels the corresponding deviation from the analytic result
884:            is amplified by $10^5$ and $10^6$, respectively.
885:            The spatial coordinate $w$ is defined in Eq.\,(\ref{w}).}
886:   \label{CCMntg}
887: \end{figure}
888: %
889: analytic values obtained for $a=1$ and a Courant factor of 0.45. 
890: In this solution no steep gradients are present
891: and 300 grid points in each region are sufficient
892: to reproduce the analytic values to within a relative error of about
893: $10^{-5}$. Again longer runs do not reveal any further features as
894: the metric settles down into Minkowskian values. We conclude that the
895: code reproduces analytic solutions with one or two
896: gravitational degrees of freedom with high accuracy over the dynamically
897: relevant time intervals.
898: 
899: 
900: %========================================================================
901: \subsubsection{Time dependent convergence analysis}
902: \label{CCM_CONVERGENCE}
903: %
904: %
905: Even though the accuracy and long term stability of the code has been 
906: demonstrated in the previous sections, we still have to make sure that
907: it is also second order convergent. In particular the
908: start-up procedure described in section \ref{start_int} and the use
909: therein of the Euler scheme to calculate the auxiliary time slice
910: at $-dt/2$ might raise questions in this respect. \\
911: % We shall see that
912: %such doubts are unjustified. \\
913: For the convergence analysis we define the
914: $\ell_2$-norm of the deviation of a numerical solution $\Psi^K$ as a function 
915: of time
916: %
917: \begin{align}
918:   \Delta \Psi^K_k &= \Psi^K_k - \Psi(x_k), \\[10pt]
919:   \ell_2[\Delta\Psi^K](t) &= \sqrt{\frac{\sum_k{\left[ \Delta \Psi^K_k(t)
920:       \right]^2}}{K}}.
921:   \label{l2norm}
922: \end{align}
923: %
924: Here $\Psi_k$ is the exact and $\Psi^K_k$ the numerical value at
925: \begin{figure}[t]
926:   \centering
927:   \epsfig{file=as_ccm_conv_nu.ps, height=200pt, width=150pt, angle=-90}
928:   \epsfig{file=as_ccm_conv_tau.ps, height=200pt, width=150pt, angle=-90}
929:   \epsfig{file=as_ccm_conv_gam.ps, height=200pt, width=150pt, angle=-90}
930:   \caption{The convergence factor $\ell_2[\Psi^{300}]/\ell_2[\Psi^{600}]$ is
931:            plotted as a function of time for the variables $\nu$,
932:            $\tau$ and $\gamma$. For our second order scheme
933:            we obtain a constant convergence factor of 4 expected for
934:            doubling the grid resolution.}
935:   \label{as_ccm_conv}
936: \end{figure}
937: %
938: grid point $k$ obtained for a total of $K$ grid points. 
939: We have calculated the $\ell_2$ norm for the Xanthopoulos solution
940: of the previous section using 300 and 600 grid points 
941: in each region. In Fig.\,\ref{as_ccm_conv} we plot the
942: quotient as a function of time. Corresponding to the increase of the grid
943: resolution by a factor of 2 we expect a convergence factor of 4 for
944: the second order scheme. In spite of the use of the first order
945: Euler method for the start-up, second order convergence is clearly maintained
946: throughout the dynamically relevant evolution.
947: 
948: 
949: 
950: 
951: 
952: 
953: 
954: 
955: 
956: