0804.4222/analysis.tex
1: \section{Analysis of test applications}
2: \label{sec:testsresults}
3: 
4: Before we test and analyze our numerical method
5: in \Sectionref{sec:numresults}, we briefly introduce some background
6: for our test application
7: in \Sectionref{sec:backgroundapplication}. More details can be found
8: in our similar discussion in \cite{beyer08:TaubNUT}, where we
9: emphasize the physical and mathematical ideas and interpret the
10: results.
11: 
12: \subsection{Background of the test application}
13: \label{sec:backgroundapplication}
14: 
15: \subsubsection{Physical and mathematical background}
16: Our aim is to compute cosmological solutions of Einstein's theory of
17: relativity; in particular we are interested in the strong cosmic
18: censorship conjecture in Gowdy vacuum solutions of Einstein's field
19: equations for spatial \Sth- or \SoXSt-topologies
20: \cite{Gowdy73,Isenberg89,Chrusciel90,garfinkle1999,beyer08:TaubNUT,Stahl02}.
21: All our discussions assume vacuum and a cosmological constant
22: $\lambda$, so that Einstein's field equations (EFE) in geometric units
23: $c=1$, $G=1/(8\pi)$ read
24: \begin{equation}
25:   \label{eq:EFE}
26:   \tilde R=\lambda \tilde g.
27: \end{equation}
28: Here $\tilde g_{\mu\nu}$ is the spacetime metric, which is the
29: fundamental unknown encoding the information about the gravitational
30: field. Its Ricci tensor $\tilde R$ \cite{hawking} is a $2$nd order
31: quasi-linear expression in the metric. We will always assume four
32: spacetime dimensions, that the signature of the metric is Lorentzian
33: $(-,+,+,+)$, and that Cauchy surfaces, i.e.\ the ``surfaces of
34: constant time'', are diffeomorphic to \Sth. Furthermore, we suppose
35: $\lambda>0$. 
36: 
37: In \cite{penrose1963,penrose1979}, Penrose introduced his notion of
38: conformal compactifications. The idea is to rescale the physical
39: metric $\tilde g$ by means of a conformal factor $\Omega$, which is a
40: smooth strictly positive function on the spacetime manifold $\tilde
41: M$. This yields the so called conformal metric
42: \[g:=\Omega^2\tilde g.\] Now, loosely speaking, if it is possible to
43: attach those points to $\tilde M$, which are the limit points of
44: vanishing $\Omega$, so that the new manifold $M$ is smooth and the
45: metric $g$ can be extended as a smooth metric on $M$, then we say that
46: the original spacetime has a smooth conformal compactification. The
47: references above, but in particular \cite{Friedrich2002}, give further
48: necessary technical requirements to make this loose statement
49: rigorous. Under those conditions, the set $\Omega=0$ is a smooth
50: surface in $M$, called conformal boundary $\scri$. Physically it
51: represent ``infinity''. In \cite{Friedrich2002}, it is shown, that
52: conformal boundaries must be spacelike hypersurfaces with respect to
53: the conformal metric for all solutions of \Eqref{eq:EFE} with
54: $\lambda>0$.  One calls such solutions ``future asymptotically
55: de-Sitter'' (FAdS) \cite{DeSitter,galloway2002}, if its conformal
56: boundary has a smooth non-empty future component $\scrip$; there is
57: the analogous concept for the past time direction. In particular, the
58: de-Sitter solution \cite{hawking} is FAdS. Under these conditions,
59: $\scrip$ represents the infinite timelike future of $\tilde M$. Some
60: of the asymptotic geometric properties of FAdS solutions are discussed
61: in \cite{beyer:PhD}.
62: 
63: Friedrich introduced his \term{conformal field equations} (CFE), as
64: reviewed for instance in \cite{Friedrich2002}, in order to deal with
65: conformally compactified solutions of Einstein's field equations. In
66: these conformally invariant equations, the fundamental unknown is the
67: conformal metric $g$ and the conformal factor $\Omega$ related to the
68: physical metric $\tilde g$. The non-trivial property of these
69: equations is, that they are, first, equivalent to Einstein's field
70: equations wherever $\Omega>0$, and, second, yield regular hyperbolic
71: evolution equations even where $\Omega=0$.  Under the assumptions
72: above, the CFE allow us to formulate what we call ``$\scrip$-Cauchy
73: problem'' \cite{DeSitter}. The idea is to prescribe data for the CFE
74: on the hypersurface $\scrip$, including its manifold structure,
75: subject to certain constraints implied by the CFE. These data can then
76: be integrated into the past by means of evolution equations implied by
77: the CFE. Friedrich proved that the \scrip-Cauchy problem is
78: well-posed, and that the unique FAdS solution corresponding to a given
79: choice of smooth data on \scrip is smooth, as long as it can be
80: extended into the past.  It is remarkable that this Cauchy problem
81: allows to control the future asymptotics of the solutions explicitly
82: by the choice of the data on \scrip. Concerning the past behavior of
83: the solution corresponding to a given choice of data on \scrip,
84: however, there is only limited understanding and a-priori control,
85: because of the complexity of the field equations.  In this paper, we
86: will give no details on the constraints on \scrip, and say only
87: briefly what the relevant initial data components are, since we do not
88: want to introduce all necessary geometric concepts now. However, we
89: write down a special class of solutions of the constraints
90: in \Sectionref{sec:id2}. We refer to \cite{DeSitter,beyer:PhD}, where
91: the details have been carried out.
92: 
93: 
94: We decided to use the so-called \term{general conformal field
95:   equations}, which are the CFE in conformal Gauss gauge
96: \cite{AntiDeSitter,Friedrich2002}.  In our applications, we specialize
97: the gauge even further to what we call \term{Levi-Civita conformal
98:   Gauss gauge} \cite{beyer:PhD}. In this paper here, we will discuss
99: neither the physical properties, nor the possibly bad implications of
100: this choice of gauge, but just refer to
101: \cite{beyer:PhD,beyer08:TaubNUT}. In any case, assuming, without loss
102: of generality, $\lambda=3$, and having fixed the residual gauge
103: initial data, as described in \cite{beyer:PhD}, the implied set of
104: evolution equations is
105: \begin{subequations}%
106:   \label{eq:gcfe_levi_cevita_evolution}%
107:   \begin{align}
108:     \label{eq:evolution_frame}
109:     \partial_t e\indices{_a^c}&=-\chi\indices{_a^b}e\indices{_b^c},\\
110:     \partial_t\chi_{ab}
111:     &=-\chi\indices{_{a}^c}\chi_{cb}
112:     -\Omega E_{ab}
113:     +L\indices{_{{a}}_{b}},\\
114:     \partial_t\Connection abc
115:     &=-\chi\indices{_a^d}\Connection dbc
116:     +\Omega B_{ad}\epsilon\indices{^b_c^d},\\
117:     \partial_t L_{ab}
118:     &=-\partial_t\Omega\, E_{ab}-\chi\indices{_a^c}L_{cb},\\
119:     \label{eq:Bianchi1}
120:     \partial_t E_{fe}-D_{e_c}B_{a(f}\epsilon\indices{^a^c_{e)}}
121:     &=-2\chi\indices{_c^c}E_{fe}
122:     +3\chi\indices{_{(e}^c}E_{f)c}
123:     -\chi\indices{_c^b}E\indices{_b^c}g_{ef},\\
124:     \label{eq:Bianchi2}
125:     \partial_t B_{fe}+D_{e_c}E_{a(f}\epsilon\indices{^a^c_{e)}}
126:     &=-2\chi\indices{_c^c}B_{fe}
127:     +3\chi\indices{_{(e}^c}B_{f)c}
128:     -\chi\indices{_c^b}B\indices{_b^c}g_{ef},\\
129:     \label{eq:conffactor}
130:     \Omega(t)&=\frac 12\, t\, (2-t),
131:   \end{align}
132:   for the unknowns 
133:   \begin{equation}
134:     u=\left(e\indices{_a^b}, \chi_{ab}, \Connection abc, L_{ab}, E_{fe},
135:       B_{fe}\right).
136:   \end{equation}
137: \end{subequations}
138: The unknowns are the spatial components $e\indices{_a^b}$ of a smooth
139: frame field $\{e_i\}$ as in \Eqref{eq:orth_frame_Ya}, with
140: $e_0=\partial_t$, which is orthonormal with respect to the conformal
141: metric, the spatial frame components of the second fundamental form
142: $\chi_{ab}$ of the $t=const$-hypersurfaces with respect to $e_0$, the
143: spatial connection coefficients $\Connection abc$, given by
144: $\Connection abc e_b=\nabla_{e_a}e_c-\chi_{ac}e_0$ where $\nabla$ is
145: the Levi-Civita covariant derivative operator of the conformal metric,
146: the spatial frame components of the Schouton tensor $L_{ab}$, which is
147: related to the Ricci tensor of the conformal metric by
148: \begin{equation*}
149: %\label{eq:defSchouton}
150: L_{ij}=R_{ij}/2
151:   -g_{ij}g^{kl}R_{kl}/12,
152: \end{equation*}
153: and the spatial frame components of the electric and magnetic parts of
154: the rescaled conformal Weyl tensor $E_{ab}$ and $B_{ab}$
155: \cite{Friedrich2002,FriedrichNagy}, defined with respect to $e_0$.  In
156: this special conformal Gauss gauge, the timelike frame field $e_0$ is
157: hypersurface orthogonal, i.e.\ $(\chi_{ab})$ is a symmetric matrix. In
158: order to avoid confusions, we point out that, in principle, the
159: conformal factor $\Omega$ is part of the unknowns in Friedrich's
160: formulation of the CFE. However, for vacuum with arbitrary $\lambda$,
161: it is possible to integrate its evolution equation in any conformal
162: Gauss gauge explicitly \cite{AntiDeSitter}, so that $\Omega$ takes the
163: explicit form \Eqref{eq:conffactor} in our gauge.  We note,
164: furthermore, that, since $(E_{ab})$ and $(B_{ab})$ are tracefree by
165: definition, we can get rid of one of the components for each of the
166: two. Our simple minded choice is the $33$-component by
167: $E_{33}=-E_{11}-E_{22}$; the same for the magnetic part.  The
168: evolution equations \Eqsref{eq:Bianchi1} and \eqref{eq:Bianchi2} of
169: $E_{ab}$ and $B_{ab}$ are derived from the Bianchi system
170: \cite{Friedrich2002}. In our gauge, the constraint equations implied
171: by the Bianchi system take the form
172: \begin{equation}
173:   \label{eq:bianchi_constraints}
174:   D_{e_c} E\indices{^c_e}
175:   -\epsilon\indices{^a^b_e}B_{da}\chi\indices{_b^d}=0,\quad
176:   D_{e_c} B\indices{^c_e}
177:   +\epsilon\indices{^a^b_e}E_{da}\chi\indices{_b^d}=0.
178: \end{equation}
179: Here, $\epsilon\indices{_a_b_c}$ is the totally antisymmetric symbol
180: with $\epsilon\indices{_1_2_3}=1$, and indices are shifted by means of
181: the conformal metric.  The other constraints of the full system above
182: are equally important, but are ignored for the presentation here.
183: Further discussions of that evolution system and the quantities
184: involved can be found in the references above.  Note that in
185: \Eqsref{eq:Bianchi1}, \eqref{eq:Bianchi2} and
186: \eqref{eq:bianchi_constraints}, the fields $\{e_a\}$ are henceforth
187: considered as differential operators, using \Eqref{eq:orth_frame_Ya}
188: and writing the fields $\{Y_a\}$ as differential operators in the
189: Euler coordinate basis as in
190: \Eqsref{eq:coordinate_repr_standard_frameY}. Seen as a system of
191: partial differential equations, the system
192: \Eqsref{eq:gcfe_levi_cevita_evolution} is symmetric hyperbolic, and
193: hence the initial value problem is well-posed.
194: 
195: Note that in this gauge, our initial hypersurface \scrip corresponds
196: to $t=0$. The past conformal boundary, if it exists, corresponds to
197: $t=2$. Hence, our time coordinate runs backwards with respect to
198: physical time.
199: 
200: These equations hold without any symmetry assumptions. In the
201: following we will assume that all unknown fields involved are Gowdy
202: symmetric. For the \oplo-code, we need to derive the evolution
203: equations of the matrix $(T\indices{_a^{a'}})$. The fact that the
204: conformal metric $g$ is $Y_3$-invariant, implies, after straight
205: forward computations, that the matrix
206: $(T_{ab}):=(T\indices{_a^{a'}}g_{a'b})$ is antisymmetric. Our choice
207: of frame transport is parallel transport with respect to the conformal
208: metric. This implies, after some algebra, that
209: \[\partial_t(T\indices{_a^{a'}})=0.\]
210: Hence, since the original system of equations is symmetric hyperbolic,
211: also the ``indirect reduction'' to \oplo-dimensions is symmetric
212: hyperbolic. So, the initial value problem for these equations is also
213: well-posed.
214: 
215: 
216: \subsubsection{A class of initial data}
217: \label{sec:id2}
218: As initial data on \scrip, we use the ``Berger data'', which are
219: solutions of the constraints derived for \U- and Gowdy symmetry in
220: \cite{beyer:PhD}. Those data are close to data of the
221: $\lambda$-Taub-NUT solutions and hence are particularly interesting
222: for the strong cosmic censorship conjecture
223: \cite{beyer08:TaubNUT}. Here, we restrict to Gowdy symmetry. Under the
224: conventions above, these data take the form
225: \begin{subequations}
226: \label{eq:ID}
227: \begin{align}
228:   \label{eq:IDframe}
229:   (e\indices{_a^b})&=\mathrm{diag}(1,1,a_3),\\
230:   (\chi_{ab})&=\mathrm{diag}(-1,-1,-1),\\
231:   \Gamma\indices{_1^1_2}&=0,\quad
232:   \Gamma\indices{_1^2_3}=-1/a_3,\quad
233:   \Gamma\indices{_2^1_2}=0, \quad
234:   \Gamma\indices{_2^1_3}=1/a_3,\\
235:   \Gamma\indices{_2^2_3}&=0,\quad
236:   \Gamma\indices{_3^1_2}=1/a_3-2a_3,\quad
237:   \Gamma\indices{_3^1_3}=0,\quad
238:   \Gamma\indices{_3^2_3}=0,\\
239:   (L_{ab})&=\mathrm{diag}\Bigl((5-3/a_3^2)/2,\,\,(5-3/a_3^2)/2,\,\,
240:   (-3+5/a_3^2)/2\Bigr),\\
241:   (B_{ab})&=\mathrm{diag}\Bigl(-4(1-a_3^2)/a_3^3,
242:   \,\,-4(1-a_3^2)/a_3^3,\,\,8(1-a_3^2)/a_3^3\Bigr),\\
243:   (E_{ab})&=\left(
244:     \begin{array}{ccc}
245:       E_0+C_2\,w_{20} & 0 & -\sqrt{2}\,a_3\, C_2\,\Re w_{21}\\
246:       0 & E_0+C_2\,w_{20} & -\sqrt{2}\,a_3\, C_2\,\Im w_{21}\\
247:       -\sqrt{2}\,a_3\, C_2\,\Re w_{21} 
248:       & -\sqrt{2}\,a_3\, C_2\,\Im w_{21} & -2(E_0+C_2\,w_{20})
249:     \end{array}\right).
250: \end{align}
251: \end{subequations}
252: The induced conformal $3$-metric of \scrip is a Berger sphere with a
253: free parameter $a_3>0$. The only inhomogeneous, i.e.\ space dependent
254: part of the initial data is given by the components $E_{ab}$.  For our
255: definition of the functions $w_{np}$, consult \cite{beyer:PhD}; we
256: just note that, with respect to the Euler coordinates, we have
257: \[w_{20}=\cos \chi,\quad w_{21}=\sin \chi\, e^{-i\rho_1}/\sqrt{2}.\]
258: For all these data, one finds
259: \begin{equation*}
260:   %\label{eq:specialT2}
261:   (T\indices{_a^{a'}})=
262:   \begin{pmatrix}
263:     0 & 2 & 0\\
264:     -2& 0 & 0\\
265:     0 & 0 & 0
266:   \end{pmatrix}.
267: \end{equation*}
268: In total this family of solutions of the constraints has three free
269: parameters $a_3>0$, $E_0\in\R$ and $C_2\in\R$. We remark that the
270: reason for the strange names of these parameters is consistency with
271: our notation in \cite{beyer:PhD}.
272: 
273: 
274: \subsubsection{Boundary control for the
275:   \texorpdfstring{\oplo}{1+1}-code}
276: \label{sec:CFEBoundary}
277: In \Sectionref{sec:practissuesS3}, we have motivated our boundary
278: control approach for the \oplo-code. Because the analysis depends
279: strongly on the particular equations and choice of frame transport, it
280: was not possible to give a further discussion there in full
281: generality. Hence, let us continue here for our special choice of
282: equations and frame transport. Due to what was said before, we have
283: \begin{equation*}
284:   %\label{eq:specialT2}
285:   (T\indices{_a^{a'}})=
286:   \begin{pmatrix}
287:     0 & 2 & 0\\
288:     -2& 0 & 0\\
289:     0 & 0 & 0
290:   \end{pmatrix},
291: \end{equation*}
292: for all $t$, for our choice of initial data. In this case, we say,
293: that the frame is ``boundary adapted''.  Now, we introduce the new
294: fields
295: \begin{equation}
296:   \label{eq:newfields}
297:   E_{11}^*:=(E_{11}+E_{22})/2, \quad E_{22}^*:=(E_{11}-E_{22})/2,
298: \end{equation}
299: and similar for the magnetic part $B_{ab}$, so that the boundary
300: system, introduced in \Sectionref{sec:practissuesS3}, yields the
301: following conditions at $\chi=0$ and $\chi=\pi$,
302: \begin{equation}
303:   \label{eq:BCs}
304:   E_{12}=E_{13}=E_{22}^*=E_{23}=0,\quad B_{12}=B_{13}=B_{22}^*=B_{23}=0,
305: \end{equation}
306: whereas $E_{11}^*$ and $B_{11}^*$ are free. For all other symmetric
307: invariant $2$-tensor fields, for instance the $2$nd fundamental form,
308: we get analogous relations, but, in addition, the 3-3-components are
309: free, if the tensor is not tracefree. Since the behavior of the
310: connection coefficients $\Gamma\indices{_a^b_c}$ can be derived on the
311: symmetry axes as well, which are the only non-tensorial objects in our
312: set of unknowns, we obtain a complete set of boundary conditions for
313: all the unknowns. The quantity $\normbc$ is now defined as the sum of
314: the actual absolute numerical boundary values of all those unknowns,
315: which are supposed to be zero there according to these
316: results. Monitoring $\normbc$ in a numerical computation, yields
317: information on how well the boundary conditions are satisfied.
318: 
319: In order to implement the \oplo-code numerically, we write the
320: unknowns in terms of the new electric and magnetic fields defined in
321: \Eqref{eq:newfields}. Actually, it would be better to introduce the
322: analogous combinations of fields for the other unknowns, but this has
323: not yet been done. Hence, so far, the code lacks a clean way of
324: enforcing e.g.\ the boundary condition $\chi_{11}-\chi_{22}=0$ at
325: $\chi=0$ and $\pi$. To circumvent this problem temporarily, we have
326: decided to work with a ``partial enforcement'' scheme, which, at a
327: given time of the evolution, enforces all boundary conditions except
328: for those of this type. In addition, we monitor the quantity
329: $\normbc$, and so far this treatment has turned out to be sufficient.
330: 
331: \subsection{Numerical results for the test application}
332: \label{sec:numresults}
333: 
334: \psfrag{t}[tl][tl][0.7]{$t$}
335: \psfrag{-ln(0.6952453959-t)}[][][0.7]{$-\ln(0.6952453959-t)$}
336: \psfrag{diff 1+1 and 2+1}[mc][tc][0.7]{$\normdiff$}
337: \psfrag{Viol BC}[mc][tc][0.7]{$\normbc$}
338: \psfrag{norm adapt}[mc][tc][0.7]{$\normadapt$}
339: \psfrag{Einstein norm}[mc][tc][0.7]{$\normeinstein$}
340: \psfrag{Bianchi constraint norm}[mc][tc][0.7]{$\normconstr$}
341: \psfrag{L1(Kretschmann)}[][][0.7]%
342: {$\left\|\mathrm{Kretschmann}-24\right\|_{L_1}/16$}
343: \psfrag{Kretschmann}[][][0.7]%
344: {$\left\|\mathrm{Kretschmann}-24\right\|_{L_1}/16$}
345: 
346: In \cite{beyer:PhD}, we have performed a couple of tests with the
347: \tplo-code, discussed the findings and drew conclusions about the
348: numerical method. Here, we rather focus on the \oplo-code, and show so
349: far unpublished tests and discussions
350: in \Sectionref{sec:analysis1+1}. Afterwards
351: in \Sectionref{sec:directcomp}, we also compare a simulation done with
352: the \tplo-code and the \oplo-code directly.
353: 
354: We just note that we have not made systematic investigations of the
355: CFL condition for our codes yet. 
356: 
357: 
358: \subsubsection{Analysis of computations with the
359:   \texorpdfstring{\oplo}{1+1}-code}
360: \label{sec:analysis1+1}
361: For our numerical test case here, we choose $a_3=0.7$, $C_2=0.1$ and
362: $E_0=0$ in \Eqsref{eq:ID} as initial data parameters, corresponding to
363: the ``large inhomogeneity case'' in \cite{beyer:PhD} and to one of the
364: simulations presented in \cite{beyer08:TaubNUT}. The associated
365: solution turns out to develop a singularity, and hence can be seen as
366: an interesting test case for our code. The evolution of a spatial norm
367: of the curvature invariant called Kretschmann scalar is shown in
368: \Figref{fig:solution}.  All the results we show here were done without
369: the automatic spatial adaption approach described
370: in \Sectionref{sec:numinfrastructurenew}, because, in order to study
371: convergence, it seems more useful to control and adapt the spatial
372: resolution manually. The adaption norm, computed with respect to
373: $E_{13}$, was used only for estimating the aliasing error. The time
374: integration was done with the adaptive $5$th order embedded RK scheme
375: with control parameters $\eta$ and $h_{\min}$ as discussed
376: in \Sectionref{sec:numinfrastructurenew}. For these runs, we decided
377: to use the ``partial enforcement'' scheme of the boundary conditions,
378: explained in \Sectionref{sec:CFEBoundary}. All runs were done with
379: double precision.
380: 
381: \begin{figure}[t]
382:   \begin{minipage}[t]{0.49\linewidth}
383:     \centering
384:     \includegraphics[width=\textwidth]{plot_singular_kretschmann}
385:     \caption{Evolution of curvature for the ``large inhomogeneity case''}
386:     \label{fig:solution}
387:   \end{minipage}\hfill
388:   \begin{minipage}[t]{0.49\linewidth}
389:     \centering
390:     \includegraphics[width=\textwidth]{s3code_constraints_early}
391:     \caption{Constraint violations at early times}
392:     \label{fig:constr_early}
393:   \end{minipage}
394: \end{figure}
395: 
396: The constraints \Eqref{eq:bianchi_constraints} are satisfied initially
397: up to machine precision. However, due to numerical errors, those
398: constraints typically become violated more and more with increasing
399: evolution time. Let us define $\normconstr$ as the $L^1$-norm of the
400: sum of the absolute values of each of the six components of the left
401: hand sides of \Eqsref{eq:bianchi_constraints} at a given instant of
402: time, all that divided by $\tr(\chi_{ab})$, in order to factor out the
403: observed collapse of the solution. $L^p$-norms of functions on \Sth
404: are always evaluated here by means of their corresponding functions on
405: \T and the standard $L^p$-norm on $\T$. Another norm, which measures
406: how well the numerical solution satisfies Einstein's field equations,
407: is
408: \[\normeinstein
409: :=\left\|(\tilde R_{ij} -\lambda \tilde
410:   g_{ij})/\Omega(t)\right\|_{L^{1}(\Sth)},\] where the Ricci tensor
411: $\tilde R_{ij}$ of the physical metric $\tilde g_{ij}$ is evaluated
412: algebraically from the conformal Schouton tensor $L_{ij}$ and
413: derivatives of the conformal factor $\Omega$. The indices in this
414: expression are defined with respect to the physical orthonormal frame
415: given by $\tilde e_i=\Omega e_i$. The norm is computed by summing over
416: the $L^1$-norms of each component.
417: 
418: Now, we will distinguish two phases of the evolution for these initial
419: data, in which different aspects and effects are important: the early
420: evolution close to \scrip for $t$ between $0.0$ and $0.69$, and the
421: late evolution close to the singularity, which we find at
422: $t\approx0.69520493$. In order to avoid confusions, we recall that the
423: terms ``early'' and ``late'' are always understood with respect to the
424: time coordinate $t$, which, however, runs backwards with respect to
425: the physical time.
426: 
427: At early times, it is achieved easily that the spatial discretization
428: error is not significant, until some later time when small spatial
429: structure starts to form more rapidly. One hint that this is true, as
430: we do not show here, is that $\normadapt$ is more or less constant
431: over a long time period, and small. Another hint is that the behaviors
432: of $\normconstr$ and $\normeinstein$ are not strongly influenced by
433: the spatial resolution. See \Figref{fig:constr_early} and
434: \ref{fig:einstein_early}, where $N$ represents the number of spatial
435: grid points, which is constant in this early regime. Indeed, the
436: higher $N$, the larger is the initial value $\normconstr$ due to
437: higher round-off errors for computing spatial derivatives. This is not
438: visible for $\normeinstein$, since this quantity is defined purely
439: algebraically in the unknowns.  In \Figref{fig:constr_early}, we see
440: that $\normconstr$ grows less, the higher the time resolution is,
441: i.e., in particular, the smaller the parameters $\eta$ and eventually
442: also $h_{min}$ are. However, we always observe at least weak
443: approximately exponential growth. In \Figref{fig:einstein_early}, we
444: see a similar behavior for $\normeinstein$. We do not show here that
445: there is neither a particular growth of $\normconstr$, nor of
446: $\normeinstein$, at the symmetry axes. Rather, the maximal growth
447: takes place, where the curvature increases most strongly. This can be
448: seen as a confirmation that our treatment of the coordinate
449: singularities works well, cf.\ \cite{beyer:PhD}. Note that there is an
450: optimal time resolution, in the sense that, if we choose a higher
451: resolution, the constraint error and $\normeinstein$ are actually
452: increased caused by higher round-off errors.  Although we do not show
453: any plots, we want to mention, that we have experimented with ``quad
454: precision''. For the \oplo-code, this yields reasonable performance
455: and has several consequences. First, the initial data for the
456: constraint violations are decreased by many orders of magnitude, since
457: those are determined primarily by the precision of the numerical
458: number representation. By choosing appropriate resolutions, we find
459: that the constraint violations and $\normeinstein$ can then be kept
460: several orders of magnitude smaller than in the double precision case
461: during the whole run. However, they always show exponential growth,
462: which suggests that this is the typical behavior of the constraint
463: propagation in our system of equations. Furthermore, quad precision
464: allows us to work in a regime in which discretization errors are much
465: larger than round-off errors, and hence it is easier to interpret
466: convergence tests.
467: \begin{figure}[t]
468:   \begin{minipage}[t]{0.49\linewidth}
469:     \centering
470:     \includegraphics[width=\textwidth]{s3code_einstein_early}
471:     \caption{Violations of EFE at early times}
472:     \label{fig:einstein_early}
473:   \end{minipage}\hfill
474:   \begin{minipage}[t]{0.49\linewidth}
475:     \centering
476:     \includegraphics[width=\textwidth]{s3code_violbc_early}
477:   \caption{Violation of boundary conditions at early times}
478:   \label{fig:violbc_early}
479:   \end{minipage}
480: \end{figure}
481: \Figref{fig:violbc_early} shows the behavior of $\normbc$. The errors
482: at the boundaries are small and stable, despite of some weak growth,
483: which is expected in situations close to a singularity. In these
484: tests, this error does not decrease for higher resolutions, which is a
485: hint that round-off errors have a significant effect. With quad
486: precision, we were able to confirm that the violations of the boundary
487: conditions become smaller, consistent with increasing resolution.
488: 
489: 
490: 
491: 
492: \begin{figure}[t]
493:   \begin{minipage}[t]{0.49\linewidth}
494:     \centering
495:     \includegraphics[width=\textwidth]{s3code_adaptnorm_late}
496:     \caption{Adaption norm at late times}
497:     \label{fig:adapt_late}
498:   \end{minipage}\hfill
499:   \begin{minipage}[t]{0.49\linewidth}
500:     \centering
501:     \includegraphics[width=\textwidth]{s3code_constraints_late}
502:     \caption{Constraint violations at late times}
503:     \label{fig:constr_late}
504:   \end{minipage}
505: \end{figure}
506: Concerning the late evolution, the following is found. The following
507: figures, \Figref{fig:adapt_late} to \Figref{fig:einstein_late}, show a
508: very small time neighborhood of the final time, where the runs were
509: stopped and where the solution blows up.  In the runs underlying these
510: plots, we adapted the spatial resolution several times during the runs
511: manually; the number $N$ in the figures is the final spatial
512: resolution in each case. Each manual spatial adaption step is visible
513: in the plots as a jump, because for different spatial resolutions, the
514: numerical values of the norms slightly change. In all these runs, we
515: choose $\eta=10^{-13}$, and hence the time steps $h$ decrease so
516: strongly that $h=h_{min}$ at that time when the runs were stopped.  In
517: this late time regime, the errors are dominated by spatial
518: discretization errors, because the solution has the property that
519: spatial structures shrink without bound.  This can be seen by looking
520: at the late time plot of the adaption norm in \Figref{fig:adapt_late}.
521: It shows, how strongly the demand for spatial resolution grows with
522: time, but also, that it is possible to gain control by increasing the
523: resolution at least temporarily. However, the demand for spatial
524: resolution increases very strongly with time, and it turns into a
525: difficult numerical issue to keep track of that eventually.  In
526: \Figref{fig:constr_late}, we demonstrate, how the choice of spatial
527: resolution influences the propagation of the constraint violations,
528: and that this quantity converges to a weakly exponentially growing for
529: sufficiently high spatial resolutions.  This is a promising result,
530: and shows that the constraint propagation is more or less under
531: control, as long as it is possible to increase the resolution in
532: practice.
533: \begin{figure}[t]
534:   \begin{minipage}[t]{0.49\linewidth}
535:     \centering
536:     \includegraphics[width=\textwidth]{s3code_violbc_late}
537:     \caption{Violation of boundary conditions at late times}
538:     \label{fig:violbc_late}
539:   \end{minipage}\hfill
540:   \begin{minipage}[t]{0.49\linewidth}
541:     \centering
542:     \includegraphics[width=\textwidth]{s3code_einstein_late}
543:     \caption{Violations of EFE at late times}
544:     \label{fig:einstein_late}
545:   \end{minipage}
546: \end{figure}
547: \Figref{fig:violbc_late} shows the violations of the boundary
548: conditions. They turn out to be very small and under control for
549: sufficiently large resolutions. The higher the final spatial
550: resolution is, the smaller these violations appear.  Finally, in
551: \Figref{fig:einstein_late} we show $\normeinstein$. We do not observe
552: a very strong difference between the various resolutions; indeed, in
553: order to make the differences visible at all, the time axis represents
554: an even smaller time neighborhood now. It is unexpected that at very
555: late times this norm is not necessarily smaller the higher the
556: resolution is, and it has to be investigated whether this is a problem.
557: 
558: \subsubsection{Comparison of a computation with the
559:   \texorpdfstring{\tplo}{2+1}- and \texorpdfstring{\oplo}{1+1}-code
560:   for a Gowdy symmetric solution}
561: \label{sec:directcomp}
562: 
563: \begin{figure}[t]
564:   \centering
565:   \includegraphics[width=0.49\textwidth]{regular_kretschmann}
566:   \caption{Evolution of curvature for the ``regular case''}
567:   \label{fig:solution_regular}
568: \end{figure}
569: 
570: \begin{figure}[t]
571:   \begin{minipage}[t]{0.48\linewidth}
572:     \centering
573:     \includegraphics[width=\textwidth]{plot_regular_violBC}
574:     \caption{Violations of the boundary conditions of the \oplo-code}
575:     \label{fig:violBCreg}
576:   \end{minipage}\hfill
577:   \begin{minipage}[t]{0.48\linewidth}
578:     \centering
579:     \includegraphics[width=\textwidth]{plot_regular_compare_2+1_1+1}
580:     \caption{Comparison of the \tplo- and \oplo-code}
581:     \label{fig:comp2+1_1+1reg}
582:   \end{minipage}  
583: \end{figure}
584: Next, we want to compare directly the numerical results obtained with
585: the \tplo-code and the \oplo-code for a Gowdy symmetric solution of
586: our equations. We restrict to a singularity free case, since we do not
587: want to be spoiled by the lack of spatial resolution, in particular
588: for the \tplo-code. The following initial data parameters are chosen:
589: $a_3=0.93$, $E_0=0$ and $C_2=0.5$, which correspond to the ``regular
590: case'' in \cite{beyer:PhD}. The corresponding solution is smooth for
591: all $0\le t\le 2$ and hence develops a smooth past conformal boundary
592: at $t=2$. \Figref{fig:solution_regular} shows the evolution of the
593: curvature invariant Kretschmann scalar. We show neither convergence
594: plots, nor the error quantities before, since the situation is very
595: similar to the early phase of the singular solution in the previous
596: section. Instead, we report only on one run done with one fixed
597: resolution: the size of the time step is $h=5\cdot 10^{-4}$ and the
598: number of spatial points is $N=40$ for the \oplo-code and $N_\chi=41$,
599: $N_{\rho_1}=21$ for the \tplo-code. We use the non-adaptive $4$th
600: order RK time integrator, and the automatic spatial adaption has been
601: switched off here as well.
602: 
603: The run with the \oplo-code was done without enforcing the boundary
604: conditions now, cf.\ \Figref{fig:violBCreg}. We see that this error
605: stays very small and is stable. In order to illustrate, how well the
606: results obtained with the \oplo- and the \tplo-code coincide, let us
607: define the following norm
608: \[\normdiff:=\Bigl|(E_{11}^{(\oplo)}|_{\chi=\pi})
609: -(E_{11}^{(\tplo)}|_{\chi=\pi,\,\rho_1=0})\Bigr|.\] Consider
610: \Figref{fig:comp2+1_1+1reg} where this norm is plotted vs.\ time.  We
611: find very good agreement between the two codes. Some deviations can be
612: expected, since in the \tplo-code, Gowdy symmetry is valid only
613: approximately; see more comments in \cite{beyer:PhD}. So far, we have
614: made no effort to explain the oscillatory behavior in both figures,
615: but we conjecture them to be caused by aliasing, since its amplitude
616: becomes smaller, the higher the spatial resolution is. We have also
617: compared more variables at other grid points and found similar
618: results.
619: 
620: 
621: %%% Local Variables: 
622: %%% mode: latex
623: %%% TeX-master: "paper"
624: %%% End: 
625: