1: \part{Applications and their analysis}
2: \label{part:analysis}
3: \chapter{Introduction}
4: In the previous \Partref{part:treatment} we have introduced our
5: underlying questions and problems and developed our method.
6: In this following part of the thesis we start by testing our method
7: in \Chapterref{ch:numexperiments}; this involves checks for formal
8: errors in the implementation but also experiments to see how well our
9: approach for the coordinate singularity on $\S$ behaves. For this we
10: compare the various possibilities to do this pseudospectrally
11: mentioned in \Sectionref{sec:discussiontancot}. In particular, these
12: tests involve computations with explicitly known solutions of the
13: linearized system but also with fully non-linear
14: regular $\lambda$-Gowdy spacetimes. In \Chapterref{ch:singulargowdy}
15: we compute singular $\lambda$-Gowdy spacetimes. Here, the emphasis
16: lies again on the study of the behavior of the code; but these
17: considerations can already be seen as first preliminary, though non-systematic,
18: investigations of some of our underlying questions. In
19: \Sectionref{sec:comparisonGowdymethods} we discuss the properties and
20: expectations for our method in the light of the
21: other numerical methods for singular Gowdy spacetimes that exist in the
22: literature. Then in
23: \Sectionref{sec:outluckfuture} and \ref{sec:summary}, we summarize,
24: name the open problems
25: and collect a
26: list of projects for future research.
27:
28: In fact, one should note that almost all applications which we present
29: here
30: belong to the Gowdy class. On the one hand the problem often
31: simplifies technically under this symmetry assumption; in particular
32: the symmetry reduces the problem to a lower number of spatial
33: dimensions. Additionally
34: in the Gowdy class, singularities can be expected to be
35: non-oscillatory. On the other hand,
36: this class of spacetimes has been studied most extensively and one has
37: quite a good overview where the open issues lie.
38:
39: Another remark is the following. During our analysis we use the norms
40: $L^p(\S)$. Usually these are defined with respect to the standard
41: measure on $\S$. However, we define the $L^p$-norm
42: of a function $f\in C^\infty(\S)$ to be the $L^p$-norm of the
43: corresponding function $f\in\breve X\subset C^\infty(\T)$
44: (\Sectionref{sec:map_Phi}) with respect to the
45: standard measure on $\T$.
46:
47:
48:
49: \chapter{Numerical experiments}
50: \label{ch:numexperiments}
51:
52: \section{Tests with explicit solutions}
53: \label{sec:experimentslinearized}
54:
55: \setcounter{totalnumber}{1}
56:
57: \subsection{Explicit solutions of the
58: spin-2-system on the de-Sitter background}
59: \label{sec:linearized_solutions}
60: In this section we want to show results from experiments with
61: solutions with
62: $\S$-topology of the linearized general conformal field equations on the
63: de-Sitter background. The purpose is
64: to check whether our numerical algorithm to compute frame derivatives
65: on $\S$ that we summarized and deepened in
66: \Sectionref{sec:discussiontancot}, works and how well it performs.
67:
68: Consider again the general conformal field equations in Levi-Civita
69: conformal Gauß gauge \Eqsref{eq:gcfe_levi_cevita_evolution}. The
70: de-Sitter solution in this gauge takes the form \Eqref{eq:dSLCCGG}
71: with conformal factor $\Omega=\frac 12t(2-t)$ from which the unknowns
72: of this system can be derived. In particular, this solution is
73: conformally flat, i.e.\ all components of the rescaled conformal Weyl
74: tensor vanish identically.
75: A solution of the
76: Bianchi system on this given background solution can be considered as the
77: solution of the
78: linearization of the conformal field equations with respect to that
79: background solution.
80: In \cite{penroserindler}, where also the
81: original references are listed, a consistency conditions is derived
82: which is necessary such that Bianchi system on a given background $M$
83: has a solution at all. This consistency condition is satisfied on any
84: conformally flat background.
85: On a given background, the Bianchi system is
86: often called \term{spin-2-system} and the rescaled Weyl tensor
87: $W\indices{^i_j_k_l}$ \term{spin-2-field}. Related discussions and
88: further terminology can
89: be found in \cite{penroserindler}. Formally one
90: gets the linearized evolution equations by setting $\Omega=0$ and
91: $\dot\Omega=0$
92: in \Eqsref{eq:gcfe_levi_cevita_evolution} which means that the
93: ``background part'' of the equations get decoupled from the
94: spin-$2$-part and we solve for the conformally flat background and the
95: spin-$2$-field on this background simultaneously. The ``background
96: part'' consists only of ODEs and the spin-$2$-part is a linear
97: symmetric hyperbolic system (apart from the small subtleties mentioned
98: in \Sectionref{sec:LCCGG}).
99:
100: Now we briefly summarize the steps to derive explicit solutions of the
101: linearized equations
102: on the de-Sitter background. The main idea is that the spin-$2$-system
103: is conformally invariant. By a suitable conformal transformation and a
104: corresponding change in the time coordinate (cf.\ the derivation of
105: \Eqref{eq:dS_conformal}) we can bring the de-Sitter
106: solution to the static Einstein cylinder. In this gauge, the evolution
107: equations derived from the
108: spin-$2$-system have constant
109: coefficients. Expanding the unknowns in terms of the basis functions
110: $\{w^n_{ik}\}$ we obtain a linear system of ODEs which is decoupled
111: for each $n$-mode but coupled among the $i$- and $k$-modes. At least
112: for $n=0$ and $n=2$, this
113: system can be solved explicitly by diagonalizing the evolution
114: matrix; the case $n>2$ has not been considered yet.
115: We do not write down the lengthy general expressions for the solutions
116: here since they are not of further interest for us.
117:
118: In any case, it is
119: not sufficient to just solve the evolution equations, also the
120: constraints \Eqsref{eq:bianchi_constraints} of the spin-$2$-system have
121: to be satisfied. This means we
122: have to choose initial data consistent with the constraints which are
123: then satisfied for all later times because on
124: the conformally flat background the constraints propagate. Choosing
125: the initial data for the magnetic part of the spin-$2$-field to vanish,
126: the only constraint left is
127: the electric constraint
128: \Eqref{eq:bianchi_electric_constraints}. How to solve this equation
129: on a general Berger sphere, in particular on $\scrip$ of
130: de-Sitter, has been discussed in
131: \Sectionref{sec:solelectrconstraintscri}.
132:
133: \subsection{Numerical solutions of the linearized equations}
134: For our numerical experiments we choose the data of
135: \Sectionref{sec:solelectrconstraintscri} with $a_3=1$,
136: $C_3=\sqrt{2}$ which means that initially
137: \[E_{11}=2\, \text{Re}(w_{21})=-E_{33},\quad E_{13}=\sqrt{2}\, w_{20}\]
138: and all other components of the spin-$2$-field vanish. As a side
139: remark, note that these initial data are the only data in this whole thesis
140: which are not Gowdy symmetric.
141: The numerical
142: solutions that we compute are compared
143: to the corresponding exact solution of
144: the linearized equations constructed above whose $E_{11}$-component reads
145: \[E_{11}=-2\,\frac{\sin (10\arctan(t-1))}{(1-t+\frac 12 t^2)^3}\,
146: \text{Re}(w_{21}).\]
147:
148: Let us define two error norms. The
149: deviation of the numerical solution from the exact solution
150: is
151: \[\normdiffexact:=
152: \left\|E_{11}^{(\mathit{num})}-E_{11}^{(\mathit{exact})}\right\|_{L^1(\S)},\]
153: and the violation of the electric constraint
154: \Eqref{eq:bianchi_electric_constraints} is
155: \begin{equation}
156: \label{eq:electric_error_norm}
157: \normelec:=\left\|D_{e_c} E\indices{^c_e}
158: -\epsilon\indices{^a^b_e}B_{da}\chi\indices{_b^d}\right\|_{L^1(\S)}.
159: \end{equation}
160: For tensorial quantities, as in the second definition, we always
161: assume summation over all components. Note however, that
162: $\normdiffexact$ takes into
163: account only one component of the solution. In general, this should
164: not be done because one could think
165: about frames which leave special components roughly regular
166: while the other components are problematic. For the analysis of this
167: simple test case in this section, we are
168: convinced that this is sufficient, but
169: later on, we will try to avoid
170: such error norms.
171:
172: We use the methods down-to-up (referred to as ``D2U''
173: in the plots) with staggered coordinate singularity
174: and the direct multiplication method (referred to as
175: ``DirMul.'') with staggered coordinate singularity; both are
176: described in \Sectionref{sec:discussiontancot}. More
177: thorough comparisons between various pseudospectral methods are done
178: in \Sectionref{sec:comparison} in the non-linear case.
179: The runs here were done with various
180: resolutions referred to as ``lSlT'' (\textit{low space low time}),
181: ``lSmT'' (\textit{low space medium time}) etc. The specific
182: resolutions are given below.
183:
184: \begin{figure}[tb]
185: \centering
186: \subfloat[$\normdiffexact$ absolute (mT)]{
187: \includegraphics[width=0.49\linewidth]{plot_linearized_diffexact_all_psfrag}}%
188: \subfloat[$\normdiffexact$ convergence]{
189: \includegraphics[width=0.49\linewidth]
190: {plot_linearized_diffexact_convergence_psfrag}}
191: \caption{Deviation from the exact solution}
192: \label{fig:deviation_linearized}
193: \end{figure}
194: Consider \Figref{fig:deviation_linearized} where we plot the deviation
195: of the numerical from the exact solution for various
196: resolutions and for the two different methods mentioned above. The
197: abscissa represents time $t$ with $t=0$ corresponding to $\scrip$ and
198: $t=2$ to $\scrim$
199: while the ordinate shows $\normdiffexact$. The left plot is devoted to
200: show these errors for a given time resolution mT, namely with time step
201: $h=5.0\cdot 10^{-4}$, but for varying spatial resolutions lS
202: $N_1=9$, $N_2=5$, mS $N_1=25$, $N_2=13$, hS $N_1=77$, $N_2=39$. Here
203: $N_1$ is the number of collocation points in the $\chi$-direction and
204: $N_2$ the number in the $\rho_1$-direction.
205: Note that the ``spiky features'' in the plot just represent the
206: oscillatority behavior of the exact solution.
207: In the right plot we show
208: convergence for varying time resolutions; here in addition to mT above
209: we have lT $h=1.0\cdot 10^{-3}$ and hT $h=2.5\cdot 10^{-4}$. In both
210: plots it is obvious that the down-to-up method works very well. The
211: absolute agreement with the exact solution is of the
212: order $10^{-10}$. The left plot implies that there is basically no
213: difference when the spatial resolution is changed which is clear since the
214: solution consists only of basis functions $\{w_{np}\}$ with $n=2$ and
215: is hence, up to round-off errors,
216: represented exactly by all three spatial resolutions, cf.\
217: \Lemref{lem:fourier_wnp}. Indeed, higher
218: spatial resolutions here can only make the numerical solution worse
219: because higher round-off errors are introduced. However, the influence
220: of round-off errors is not notable yet and everything is very stable
221: including our treatment of the coordinate singularity. To drive the
222: method instable, much higher spatial resolutions have to be used, see
223: \Figref{fig:linearized_instability}. The right plot demonstrates nice
224: $4$th-order convergence
225: of the code with down-to-up in time. This means that the errors are
226: dominated by the
227: time discretization for both the low and medium (and in fact also for
228: the high) spatial resolutions and then the $4$th-order Runge-Kutta method
229: enforces $4$th-order convergence. What is
230: interesting about the left picture of
231: \Fignref{fig:deviation_linearized} is that the direct
232: multiplication method, i.e.\ the naive way of treating the coordinate
233: singularity, is strongly instable\footnote{Here, by ``instable'' we do not
234: mean the rigorous notion of \Sectionref{sec:convergence} but a
235: numerical solutions which deviates from its expected behavior and
236: blows up strongly.}. Higher resolution strengthen the
237: instability which can be explained because the spatial points get
238: closer to the coordinate
239: singularity.
240:
241: \begin{figure}[tb]
242: \begin{minipage}[t]{0.49\linewidth}
243: \centering
244: \includegraphics[width=\linewidth]{plot_linearized_diffexact_instable_psfrag}
245: \caption{Instability for too high spatial resolutions}
246: \label{fig:linearized_instability}
247: \end{minipage}
248: \begin{minipage}[t]{0.49\linewidth}
249: \centering
250: \includegraphics[width=\linewidth]{plot_linearized_constraint_convergence_psfrag}
251: \caption{Convergence of the constraint violations (lS)}
252: \label{fig:linearized_convergence_constraint}
253: \end{minipage}
254: \end{figure}
255: In \Figref{fig:linearized_convergence_constraint} we also demonstrate
256: convergence of the constraint violations of the linearized
257: solution. This is actually not so easy because the initial
258: constraint violations are of the order of the machine round-off errors
259: and it stays like that during the evolution; hence we cannot
260: expect to find a converging constraint violation since round-off
261: errors spoil convergence. To see convergence of the constraints at all
262: we introduced an artificial
263: constraint violation of the order $10^{-8}$ for the runs underlying
264: \Fignref{fig:linearized_convergence_constraint}. The constraint
265: propagation system implies that the constraint violations are
266: oscillatory in this simple case. However, even for that we find that
267: the difference in the constraint violation for the various resolutions
268: above is also of the order of the round-off error and still no
269: convergence can be observed. Hence, these runs (and only the runs for
270: this plot) are done at \textit{very} low time resolutions; here lT
271: $h=4.0\cdot 10^{-2}$, mT $h=2.0\cdot 10^{-2}$, hT $h=1.0\cdot
272: 10^{-2}$, hhT $h=0.5\cdot 10^{-2}$ while spatial resolutions are as
273: above. Then, we are able to observe as shown in the plot, that there
274: is $4$th-order convergence but only for the three lower time
275: resolutions. For the highest time resolution hhT at least the orders of
276: magnitude are still correct. For even higher resolution no convergence
277: at all would be visible.
278:
279: This already shows a problem for the interpretation of our numerical
280: results. Since in some applications pseudospectral methods are so
281: accurate that round-off errors dominate, the standard
282: convergence tests, that work very nicely for finite differencing
283: methods in particular, have to be handled with care.
284:
285: \begin{figure}[tb]
286: \centering
287: \includegraphics[width=0.49\linewidth]{plot_linearized_constraint_nonlinear_psfrag}
288: \caption{Comparison of constraint violations in the linearized and in the
289: non-linear case}
290: \label{fig:linearized_nonlinear}
291: \end{figure}
292: Finally consider \Figref{fig:linearized_nonlinear} where we compare
293: the propagation behaviors of the constraint in the linear and
294: in the non-linear case, i.e.\ when the coupling between the
295: ``background part'' and the spin-$2$-system is switched on again. We
296: do this for two different spatial resolutions but with one time
297: resolution. One can expect that the non-linear case is much more
298: severe since the non-linear coupling induces fine structure
299: formation. In particular, the medium spatial resolution is only
300: sufficient at early times in the non-linear case while for the high
301: resolution there is not very much difference between the violations in
302: the linear and the non-linear case which means that hS is sufficient
303: to resolve all fine structure formed. Hence, in generic non-linear
304: runs one can expect that the demand for spatial resolution increases
305: with time and hence it will be crucial to have some method which keeps
306: track of this demand. A possible, and so far working approach to this
307: is the spatial adaption method introduced in
308: \Sectionref{sec:spatial_adaption}. This will be used intensively in
309: the computation of singular spacetimes in
310: \Chapterref{ch:singulargowdy}. Note another typical feature from
311: \Fignref{fig:linearized_nonlinear}: the initial constraint violation
312: is higher the higher the spatial resolution is chosen initially due to
313: the influence of round-off errors. This
314: has a some impact for later discussions.
315:
316: \subsection{Other explicit solutions}
317: To further check the implementations of the equations I have also
318: reproduced successfully the explicitly known $\lambda$-Taub-NUT family
319: \Sectionref{sec:TaubNUT} in the way mentioned at the end of
320: \Sectionref{sec:solelectrconstraintscri}. We will not discuss
321: this further here.
322:
323:
324:
325: \section{Regular \texorpdfstring{$\lambda$}{lambda}-Gowdy
326: spacetimes on \texorpdfstring{$\S$}{S3}}
327: \label{sec:comparison}
328:
329: \setcounter{totalnumber}{2}
330:
331: In this section we make numerical experiments with the methods
332: introduced and discussed in \Sectionref{sec:discussiontancot} with the
333: full non-linear evolution equations, namely
334: the general conformal field equations in Levi-Civita conformal Gauß
335: gauge as described in \Sectionref{sec:LCCGG}. As initial data on
336: $\scrip$ we
337: choose a Gowdy
338: initial data set on a Berger sphere as derived
339: in \Sectionref{sec:solelectrconstraintscri} that turns out to be in
340: the de-Sitter
341: stability region dSSR (\Sectionref{sec:non-linear stability of de-Sitter}).
342: The initial data parameters are
343: \[a_3=1,\,a_3=0.92,\,(E_{11})_{0,0}=(E_{22})_{0,0}=0,\,C_2=0.5,\]
344: cf.\ \Eqsref{eq:n2solelectrconstraint}; all other $C_i$ vanish
345: because we demand Gowdy symmetry. The corresponding solution turns out
346: to be
347: both future and past asymptotically de-Sitter, but which nevertheless has
348: some non-trivial amount of inhomogeneity.
349:
350: In this section, results from five different pseudospectral methods to
351: compute frame derivatives on $\S$ (\Sectionref{sec:discussiontancot})
352: are presented and compared.
353: The first method, referred to as
354: ``D2U Stag.'' in the following plots, is the down-to-up method
355: introduced in \Sectionref{sec:discussiontancot} with staggered coordinate
356: singularities, i.e.\ the collocation points are shifted such that all
357: coordinate
358: singularities are exactly in the middle between two grid
359: points. Correspondingly, we make the same computation with the method
360: ``D2U Non-Stag.'' where the coordinate singularities coincide
361: with the relevant grid points. We also show results of the method
362: ``D2UMod Stag.'' which is just a reimplementation of the
363: down-to-up method above but which is arranged slightly
364: differently such that the distribution of round-off errors is not
365: the same. We expect that the results of this method are the same as for
366: ``D2U Stag.'' but the influence of round-off errors is hard to
367: predict; eventually it turns out that there are indeed differences,
368: see below.
369: Furthermore, we did runs with the method ``U2D Stag.'' and ``DirMul.'', cf.\
370: \Sectionref{sec:discussiontancot}. For the direct multiplication
371: method ``DirMul.'' the
372: coordinate singularities are staggered between the grid points.
373:
374: The runs were performed with various
375: resolutions referred to as ``lSlT'' (\textit{low space low time}),
376: ``lSmT'' (\textit{low space medium time}), ``lShT''
377: (\textit{low space high time}), ``mShT''
378: (\textit{medium space high time}) etc. The low resolution in time
379: corresponds to a time step $h=2\cdot 10^{-3}$, the medium one to
380: $h=1\cdot 10^{-3}$ and the high one to $h=5\cdot 10^{-4}$. Note that
381: $t=0$ corresponds to $\scrip$ and $t=2$ to $\scrim$ thus the complete
382: spacetime is computed. The
383: low resolution in space is $N_1=25$, $N_2=13$ where $N_1$ is the
384: number of collocation points in $\chi$-direction and $N_2$ the number
385: in $\rho_1$-direction. The medium resolution in space is given by
386: $N_1=41$, $N_2=32$.
387:
388:
389: It is important to note
390: that here we do not have explicit solutions to compare with;
391: other measures of the numerical errors have to be found. We use
392: the violation norm of the electric constraint
393: \Eqref{eq:electric_error_norm} and
394: the following measure of the violation of Einstein's vacuum field
395: equations \Eqref{eq:EFE}
396: \[\normeinstein
397: :=\left\|(\tilde R_{ij}-\lambda \tilde g_{ij})/
398: \Omega(t)\right\|_{W^{1,1}(\S)}.\]
399: By the $W^{1,1}$-norm of a quantity $u$ we mean the sum of the
400: $L^1$-norm of $u$ and the $L^1$-norms of all $\{Y_a\}$-derivatives of
401: $u$. For the tensorial quantities we additionally sum over all
402: components as we will do in all similar expressions in the following.
403: The physical Ricci tensor and the physical metric are
404: projected onto the physical orthonormal frame. Since this expression
405: is $O(\Omega)$ close to $\scri$ with $\Omega$ the physical conformal
406: factor, this factor is divided out. Another norm that monitors the
407: behavior of the numerical code is
408: $\normadapt$ introduced in \Sectionref{sec:spatial_adaption}; however,
409: in this section this norm is not yet used for automatic spatial adaption.
410: A further norm is
411: \[\normweyl
412: :=\left\|E_{ab}\right\|_{L^1(\S)}+\left\|B_{ab}\right\|_{L^1(\S)}\]
413: which gives us the order of magnitude of certain relevant components
414: of the solution at a given
415: time. Another error quantity measures the quality of Gowdy symmetry. As
416: explained in \Sectionref{sec:implementationOfS3Gowdy}, the symmetry
417: along $Y_3$ is not enforced during time evolution and
418: there is the possibility that the solution, although initially fully
419: Gowdy-symmetric, strongly deviates. To check this we thus define the
420: norm
421: \[\normkilling
422: :=\left\|K_{ab}\right\|_{L^1(\S)}\]
423: where the Killing operator $K_{ab}$ is defined by
424: \[K_{ab}:=g([Y_3,e_{(a}],e_{b)}),\]
425: cf.\ \Eqref{eq:killing_eq_onf}; note that
426: $[Y_3,e_0]=0=g([Y_3,e_a],e_0)$ by the choice of gauge.
427:
428: \begin{figure}[tb]
429: \centering
430: \subfloat[lSmT $N_1=25$, $N_2=13$, $h=1.0\cdot 10^{-3}$]{
431: \includegraphics[width=0.49\linewidth]{plot_compare_constraint_All_lSmT_psfrag}}%
432: \subfloat[lShT $N_1=25$, $N_2=13$, $h=0.5\cdot 10^{-3}$]{
433: \includegraphics[width=0.49\linewidth]{plot_compare_constraint_All_lShT_psfrag}}\par
434: \subfloat[mSmT $N_1=41$, $N_2=32$, $h=1.0\cdot 10^{-3}$]{
435: \includegraphics[width=0.49\linewidth]{plot_compare_constraint_All_mSmT_psfrag}}%
436: \subfloat[mShT $N_1=41$, $N_2=32$, $h=0.5\cdot 10^{-3}$]{
437: \includegraphics[width=0.49\linewidth]{plot_compare_constraint_All_mShT_psfrag}}
438: \caption{Violation of electric constraint for the methods discussed
439: in the text}
440: \label{fig:electric_viol_all_methods}
441: \end{figure}
442:
443: Consider \Figref{fig:electric_viol_all_methods}. Here we plot the
444: violation of the electric constraint vs.\ time for the methods above
445: where each
446: picture shows a fixed resolution. Exponential constraint growth, as we
447: observe here, can be
448: expected in most free evolutions so the questions
449: are rather, first, how strong this growth is, second, if the
450: violations stays small
451: enough for the relevant time interval such that the
452: solution can be trusted, and third, what the main reason for the growth
453: is. The down-to-up variants all work
454: very well and are stable. There is no difference if the coordinate
455: singularity is staggered or not as expected. For the highest
456: resolution the modified down-to-up method ``D2U Mod'' is a bit worse
457: which shows that the influence of round-off errors for different
458: implementations of the same thing can behave quite
459: differently. However, this method can also be considered as
460: stable. Definitely instable are the direct multiplication and the
461: up-to-down methods. In particular, the instability is stronger the
462: higher the spatial resolution is. For the direct
463: multiplication method this can be explained because the grid points
464: get closer to the coordinate singularity for higher resolutions. For
465: the up-to-down method, it seems to be true what we suspected already
466: in \Sectionref{sec:discussiontancot}: the round-off errors which are
467: relatively strong in high frequencies get distributed to the low
468: frequencies and this drives instability. Hence, these two methods can
469: be considered as
470: unusable. Increasing the time resolution yields a smaller constraint
471: violation at least for low times; we check for convergence
472: in \Figref{fig:electric_viol_3_methods}. As we have already seen in
473: \Figref{fig:linearized_nonlinear}
474: the errors for later times can be dominated by a lack in spatial
475: resolution, in particular for such fixed resolution runs. This results in
476: oscillations and we can observe in
477: \Fignref{fig:electric_viol_all_methods} that their amplitudes are
478: smaller the higher the spatial resolution is. Comparing the late time
479: behavior of the third and the fourth plot it can be seen that the
480: mean magnitude of the constraint violations does not
481: change although the time resolution increases. This gives us
482: another hint that error quantities in pseudospectral
483: codes must sometimes be interpreted differently than in
484: finite-differencing approaches. The problem is the following. Constraint
485: violations of the Bianchi system are propagated by means of an
486: homogeneous
487: symmetric hyperbolic system of equations, the subsidiary system
488: (\Sectionref{sec:maximaldevelopments}). In particular, if the
489: constraint violations vanish initially, they will vanish for all
490: times. But in a pseudospectral code, the initial constraint violation
491: is of the order of the machine precision, which is not zero, and hence
492: the corresponding
493: solution of the subsidiary system does not vanish. In fact,
494: depending on the properties of the evolution system, the constraints
495: do typically grow exponentially in time. Hence, from some resolution on, the
496: constraint violations cannot be decreased anymore by increasing the
497: resolution since the ``true'' solution of the subsidiary system with
498: initial data of the order of machine
499: precision gets resolved. In fact then, increasing the resolution, increases
500: the initial
501: data for the constraint quantities which yields an even higher
502: constraint
503: violations as the corresponding solution of the subsidiary system. In
504: any case, since practically we cannot increase the machine
505: precision\footnote{The machine precision can indeed be increased but
506: only with a significant loss of performance.}, the only possibility
507: that is left is to decrease the actual
508: solution of the constraint propagation system by changing the
509: properties of the evolution system. In the most
510: prominent formulations of Einstein's field equations, there are attempts
511: to introduce so called \term{constraint damping terms} into the
512: evolution system such that the corresponding constraint propagation
513: system drives the constraint violations to zero; for instance
514: \cite{Brodbeck98,Gundlach05}. However, similar techniques have not been applied
515: to the Bianchi system yet although there are analyses of the
516: constraint propagation in \cite{Frauendiener04}.
517: In any case, what we can see here is that
518: given high enough resolution, our method is able to approximate the actual
519: solution of the subsidiary system quite accurately. With
520: such resolutions the constraint
521: errors and the round-off errors, but not the discretization error,
522: dominate the errors of the solution.
523:
524: \begin{figure}[tb]
525: \centering
526: \subfloat[lS $N_1=25$, $N_2=13$]{
527: \includegraphics[width=0.49\linewidth]{plot_compare_constraint_3Main_lS_psfrag}}%
528: \subfloat[mS $N_1=41$, $N_2=32$]{
529: \includegraphics[width=0.49\linewidth]{plot_compare_constraint_3Main_mS_psfrag}}
530: \caption{Violation of electric constraint for the three stable methods}
531: \label{fig:electric_viol_3_methods}
532: \end{figure}
533: The discussion of \Figref{fig:electric_viol_3_methods} is
534: related. Here we plot again the behavior of constraint violations
535: versus time, but we restrict to the three stable methods. In each of
536: the two plots the spatial resolution is fixed while we vary the time
537: resolution. Although it is not explicitly checked in this plots, we
538: find $4$th-order convergence for low times. The same phenomena
539: observed as before, that
540: for later times the spatial discretization contributes to the errors,
541: can be observed and this results in oscillations.
542: \begin{figure}[tb]
543: \centering
544: \includegraphics[width=0.49\linewidth]{plot_compare_adapt_norm_psfrag}
545: \caption{$\normadapt$ (\Sectionref{sec:spatial_adaption}) for some
546: resolutions}
547: \label{fig:adapt_norm}
548: \end{figure}
549: Related to this is \Figref{fig:adapt_norm} where we show the behavior
550: of the adaption norm. Recall again that adaption is not used in these
551: runs; hence the apparent dynamics in this norm is caused exclusively
552: by the same
553: oscillations which we have also seen in the other norms. We see, as
554: expected, that for low spatial
555: resolutions, approximately independent of the time resolution, the
556: amount of power in the high frequencies is quite high at later
557: times and the
558: aliasing effect (\Sectionref{sec:pseudospsectr_background}) is then
559: responsible for these oscillations. For higher spatial
560: resolutions, the power in the high frequencies is correspondingly
561: smaller and hence the aliasing effect is not dominant.
562:
563: \begin{figure}[tb]
564: \centering
565: \subfloat[lS $N_1=25$, $N_2=13$]{%
566: %\label{fig:error_comparison1:a} %% hallo
567: \includegraphics[width=0.49\linewidth]{plot_compare_EinsteinNorm_All_lS_psfrag}}
568: \subfloat[mS $N_1=41$, $N_2=32$]{%
569: \includegraphics[width=0.49\linewidth]{plot_compare_EinsteinNorm_All_mS_psfrag}}
570: \caption{Violation of Einstein's vacuum equations for various
571: resolutions and methods}
572: \label{fig:einstein_all_methods}
573: \end{figure}
574: As we discussed before, the ability to use $\normelec$ as an error monitor
575: quantity has limitations. That is why we also discuss the behavior of
576: $\normeinstein$ for all the methods above in
577: \Fignref{fig:einstein_all_methods}. This error
578: quantity measures the error in the standard ADM evolution and constraint
579: equations (\Sectionref{sec:maximaldevelopments}). In fact, the
580: electric constraint corresponding to the
581: quantity $\normelec$ is only one of a large
582: number of constraints in our formulation of the field equations, and
583: it is
584: one differential order higher than the standard ones included in
585: $\normeinstein$.
586: It is difficult to understand how all these constraints somehow sum up
587: to the standard constraints. To measure these higher order errors, the
588: first spatial derivatives of the deviations from Einstein's field
589: equations are included by the $W^{1,1}$-norm. Now, have a look at
590: \Fignref{fig:einstein_all_methods}. One sees that all methods agree
591: for some time then suddenly, dependent on the spatial resolution, the
592: direct multiplication method and later also the up-to-down method
593: strongly deviate. What is surprising is that in the plots before we saw
594: the two instable methods to deviate much earlier and much stronger.
595: \begin{figure}[tb]
596: \centering
597: \includegraphics[width=0.49\linewidth]{plot_compare_ordersmagnitude_mSmT_psfrag}
598: \caption{Comparison of the orders of magnitude for the different
599: norms (mSmT)}
600: \label{fig:orders_magnitude}
601: \end{figure}
602: To obtain some deeper understanding look at
603: \Figref{fig:orders_magnitude}. There we see that the deviation of
604: $\normeinstein$ happens at that time when $\normelec$ gets higher than the
605: order of magnitude of the solution itself, represented by
606: $\normweyl$. Hence, in this sense, $\normeinstein$ is on the one hand
607: much less sensitive to errors than $\normelec$ but on the other hand
608: indicates when the errors really begin to dominate the
609: solution. This should be of general interest since the standard constraints
610: included in $\normeinstein$ are the typical error monitors in
611: numerical relativity.
612:
613: \begin{figure}[tb]
614: \begin{minipage}[t]{0.49\linewidth}
615: \centering
616: \includegraphics[width=\linewidth]
617: {plot_compare_EinsteinNorm_convergence_All_psfrag}
618: \caption{Convergence of $\normeinstein$}
619: \label{fig:einstein_convergence}
620: \end{minipage}
621: \begin{minipage}[t]{0.49\linewidth}
622: \centering
623: \includegraphics[width=\linewidth]{plot_compare_killing_psfrag}
624: \caption{Violation of Gowdy symmetry}
625: \label{fig:killing_norm}
626: \end{minipage}
627: \end{figure}
628: In \Figref{fig:einstein_convergence} we show convergence of
629: $\normeinstein$. The errors induced by time discretization seem to
630: play the dominant role for $\normeinstein$ and we find $4$th-order
631: convergence for all times.
632:
633: In \Figref{fig:killing_norm} we show the behavior of
634: $\normkilling$. We have already explained above that the code does not
635: explicitly enforce the invariance along $Y_3$, and one cannot exclude that
636: numerical errors
637: together with possible non-linear instabilities of the continuum evolution
638: equations drive the solution away from this symmetry. This plot shows
639: that this does not seem to be the case. The
640: violation of the symmetry is a bit stronger when the resolution is
641: higher but this is to some degree also caused by the higher round-off
642: errors in the spatial derivative that one has to take to compute
643: $\normkilling$. Although these investigations here cannot be
644: considered as systematic this result gives us indeed a hint that there
645: is not a strong non-linear instability of the class of Gowdy solutions
646: within the class of
647: $\U$-symmetric solutions. We come back to this later.
648:
649: Another interesting experiment, which could have been done, is to
650: give data on $\scrip$, compute, as we did here, for the
651: corresponding solution the
652: data on $\scrim$ backward in time, next use that
653: data as new initial data on $\scrim$ and finally compute the
654: corresponding data on $\scrip$ forward in time. The difference of the
655: original data
656: and the resulting data on $\scrip$, which would be zero in an exact
657: solution, can be interpreted as another
658: error measure. Further it would be interesting if those
659: oscillations, visible for instance in $\normelec$ in
660: \Figref{fig:electric_viol_3_methods} caused by a lack of spatial
661: resolution, would start close to $\scrim$ and would stop when $\scrip$ is
662: approached in the same way as we have observed here.
663:
664: In \Sectionref{sec:comparisonGowdymethods} we discuss some implications
665: of our results here
666: and of the following chapter in comparison with the existing
667: numerical methods
668: for the Gowdy class of spacetimes.
669:
670:
671:
672: \chapter{Singular \texorpdfstring{$\lambda$}{lambda}-Gowdy
673: spacetimes}
674: \label{ch:singulargowdy}
675:
676: %\section{Introduction}
677: %\label{sec:singintroduction}
678: In \Sectionref{sec:comparison} we computed regular $\lambda$-Gowdy
679: spacetimes with spatial $\S$-topology. In particular, these spacetimes
680: are geodesically complete and there is no curvature singularity.
681: Here we want to compute
682: and analyze FAdS $\lambda$-Gowdy spacetimes that are past
683: singular by modifying the initial data parameters such that the
684: de-Sitter stability region (\Sectionref{sec:dSSR}) is left. While the
685: regular class
686: excludes the $\T$-topology, see the
687: singularity theorems in \Sectionref{sec:singularitytheoremGA}, in the
688: singular class both $\T$- and $\S$-topologies are allowed and will be
689: studied here.
690:
691: As
692: the reader can
693: imagine, the singular class is much more difficult to study
694: technically than the regular class. Our results should be considered
695: to a large degree as tests of our method, both the numerical approach and the
696: formulations of the field equations, to find out the strengths and
697: limitations. However, we also present some preliminary
698: results of fundamental interest together with their investigations and
699: discussions.
700: In the first section we show numerical results for the conformal
701: field equations in the Levi-Civita
702: conformal Gauß gauge (\Sectionref{sec:LCCGG}) with $\T$-topology. Here
703: the Gowdy
704: symmetry is used to reduce the
705: evolution equations to a $1+1$-form explicitly. In the second section, similar
706: investigations are presented for the case
707: of $\S$-topology; note again that up to now the code is reduced only
708: to $2+1$ as explained before.
709: $2+1$-simulations in
710: the $\T$-case have
711: not been tried yet but are expected to behave similarly as in the
712: $\S$-case.
713:
714: In the \S-case the
715: solutions can be considered as non-linear perturbations of
716: $\lambda$-Taub-NUT spacetimes (\Sectionref{sec:TaubNUT})
717: and there are, in the special case considered, indications for an
718: interesting stability of the Taub-NUT Cauchy horizon; however, the
719: investigations have not been thorough enough yet. Further we reconsider the
720: non-linear stability issue of the class of
721: Gowdy spacetimes within the
722: class of $\U$-symmetric spacetimes, which already came up in
723: \Sectionref{sec:comparison}.
724: Afterwards,
725: we do numerical investigations with the
726: commutator field equations introduced in
727: \Sectionref{sec:commutatorfieldequations}.
728: Note again, that this
729: system is currently restricted to $\T$-topology.
730:
731: \section{Runs with the GCFE in Levi-Civita conformal Gauß gauge}
732: \label{sec:runsGCFE}
733: \subsection{Runs with \texorpdfstring{$\T$}{T3}-topology}
734: \label{sec:singularT3}
735: This section is devoted to the study of numerically generated
736: singular $\lambda$-Gowdy spacetimes with spatial
737: $\T$-topology. We make use of the initial data constructed in
738: \Sectionref{sec:T3initialdata} with the following two choices:
739: \begin{enumerate}
740: \item non-polarized data
741: \[(W_{ab})=
742: \begin{pmatrix}
743: 10^{-4} & 0 & 0\\
744: 0 & 0 & \sin x_1\\
745: \sin x_1 & 0 & -10^{-4}
746: \end{pmatrix},\]
747: \item polarized data
748: \[(W_{ab})=
749: \begin{pmatrix}
750: 10^{-4} & 0 & 0\\
751: 0 & \sin x_1 & 0\\
752: 0 & 0 & -10^{-4}-\sin x_1
753: \end{pmatrix}.\]
754: \end{enumerate}
755: As before, we assume that the other initial data quantities are given to induce
756: the Levi-Civita conformal Gauß gauge, with
757: $t=0$ the initial hypersurface $\scrip$. The time $t=2$ would correspond to
758: $\scrim$ but due to the singularity theorems in
759: \Sectionref{sec:singularitytheoremGA}, a smooth $\scrim$ cannot exist.
760:
761: %\begin{floatingfigure}{0.50\linewidth}
762: \begin{figure}[tb]
763: \begin{minipage}[t]{0.49\linewidth}
764: \centering
765: \includegraphics[width=\linewidth]
766: {plot_T3Sing_adaptnorm_psfrag}
767: \caption{$\normadapt$ for the non-polarized case}
768: \label{fig:typical_adaption_norm}
769: \end{minipage}
770: \begin{minipage}[t]{0.49\linewidth}
771: \centering
772: \includegraphics[width=\linewidth]
773: {plot_T3Sing_KretschPhys_LInfty_psfrag}
774: \caption{Kretschmann scalar for the non-po\-lar\-ized case}
775: \label{fig:KretschDiv}
776: \end{minipage}
777: \end{figure}
778: %\end{floatingfigure}
779: Let us start with the spacetime corresponding to the non-polarized
780: initial data. It turns out that at
781: $t\approx 0.9$ the variables in the equations blow up. It is a
782: curvature singularity because the Kretschmann scalar is divergent, see
783: \Figref{fig:KretschDiv}. In these runs, we employ
784: our
785: adaption techniques explained in \Sectionref{sec:spatial_adaption} as
786: one can see in \Figref{fig:typical_adaption_norm}. There, the time axis is
787: exponentially stretched and one sees that the demand for spatial resolution
788: increases almost exponentially in time. The runs were done with
789: several resolutions. The resolution, which we
790: refer to as hShT, starts from $N=11$ and
791: $h=6.25\cdot 10^{-5}$ and stops at the final time with $N=615$ and
792: $h=4.6875 \cdot 10^{-7}$ making use of the adaption methods described
793: before. For the other resolutions, the automatic
794: adaption mechanism is switched off and the adaption history of the
795: hShT run is copied, on the one hand with half of the
796: spatial resolution for the lShT run, and other other hand with half
797: the time resolution for the hSmT run etc. Here
798: $N$ is the number of collocation
799: points in $x_1$ direction. Recall that these runs are $1+1$.
800:
801: \begin{figure}[tb]
802: \centering
803: \subfloat[Behavior of $\normelec$]{
804: \includegraphics[width=0.49\linewidth]{plot_T3Sing_ConstrE_psfrag}}%
805: \subfloat[Behavior of $\normeinstein$]{
806: \includegraphics[width=0.49\linewidth]
807: {plot_T3Sing_EinsteinNorm_psfrag}}
808: \caption{Some error norms for the non-polarized case}
809: \label{fig:T3Gowdyerror}
810: \end{figure}
811: \Figref{fig:T3Gowdyerror} shows error norms for the runs for the
812: resolutions above. On the left $\normelec$ and on the
813: right $\normeinstein$ are plotted vs.\ time. From the left plot we can
814: deduce that for the high spatial
815: resolution the error is dominated by the time
816: discretization for low times as before. For larger times, increasing
817: the time
818: resolution stops to make a difference. This is so because
819: on the one hand spatial discretization plays a bigger role for later
820: times despite of the adaption; however, apparently the aliasing effect
821: is not dominant since no oscillations in the error norms can
822: be observed. On the other hand, the constraint error cannot be
823: made smaller than the actual solution of the subsidiary
824: system as we have discussed before.
825: Hence, so far, everything is in agreement with the results we found
826: before. The low spatial resolution is surely not enough and we see
827: large errors in the constraint although the code is still stable.
828: For the other runs, one
829: should note that although the constraint violation gets to the
830: order $10^{-6}$ at the final time, the relevant
831: quantities of the unknowns are of the order $10^4$ (as we do not show
832: here), and this
833: means that our accuracy is of the order $10^{-10}$. This is in
834: agreement with $\normeinstein$ at the final time. As before,
835: $\normeinstein$ is much less sensible for errors caused by spatial
836: discretization.
837:
838: \begin{figure}[tb]
839: \centering
840: \subfloat[Orbit volume density]{%
841: \label{fig:T3GowdygeometricquantitiesOVol}
842: \includegraphics[width=0.49\linewidth]{plot_T3Sing_OVol_1D_psfrag}}%
843: \subfloat[Kretschmann scalar]{%
844: \label{fig:T3GowdygeometricquantitiesKretsch}
845: \includegraphics[width=0.49\linewidth]
846: {plot_T3Sing_KretschPhys_1D_psfrag}}\par
847: \subfloat[Scalar product of the Killing vector fields]{%
848: \label{fig:T3GowdygeometricquantitiesScalarpr}
849: \includegraphics[width=0.49\linewidth]
850: {plot_T3Sing_ScalarprKV_1D_psfrag}}%
851: \subfloat[Hyperbolic velocity]{%
852: \label{fig:T3GowdygeometricquantitiesHypVel}
853: \includegraphics[width=0.49\linewidth]{plot_T3Sing_HypVel_1D_psfrag}}
854: \caption{Spatial behavior of geometric quantities for the
855: non-polarized case}
856: \label{fig:T3Gowdygeometricquantities}
857: \end{figure}
858: In \Figref{fig:T3Gowdygeometricquantities} we plot the spatial
859: distribution of some geometric
860: quantities for the singular $\T$-Gowdy solutions for three
861: different times close to the singularity. In
862: \Fignref{fig:T3GowdygeometricquantitiesOVol} we see how the
863: orbit volume density behaves. The orbit volume density is defined as the
864: square root of the determinant of the matrix
865: $(g(\partial_A,\partial_B))$ with $A,B=2,3$, i.e.\ it is, to be
866: precise, the
867: \textit{conformal} orbit volume density; but note that close to
868: $t\approx 0.9$ there is not much difference between conformal and physical
869: quantities. Here one sees one particular drawback of the Levi-Civita
870: conformal Gauß gauge compared to areal gauge. In the latter gauge, the
871: orbit volume density is a constant on each $t=const$ slice and
872: hence the singularity is approached in a ``homogeneous'' way. This is
873: not the case for our gauge here and this ``inhomogeneity'' is even increased
874: the further the singularity is approached. Some points, namely those
875: where intuitively gravity is stronger, are pulled
876: faster to the singularity than other points. In any case, such a
877: behavior can be
878: expected from a Gauß like gauge as ours. In
879: \Fignref{fig:T3GowdygeometricquantitiesKretsch}, where we plot the
880: physical Kretschmann scalar according to
881: \Eqref{eq:physKretschConfQuant}, we see that exactly at those points,
882: which approach the singularity most quickly, the Kretschmann scalar
883: blows up fastest. To avoid confusion note that the downward
884: pointing ``spiky features'' in
885: this plot are caused by the fact the I plot the absolute value of
886: the Kretschmann scalar and the ordinate is logarithmic. The plot
887: \Fignref{fig:T3GowdygeometricquantitiesScalarpr} shows
888: the scalar product of the killing vector fields, i.e.\
889: $g(\partial_2,\partial_3)$ which are related to the quantity $Q$ in
890: the standard parametrization of the Gowdy metric. Hence one sees
891: that this solution is definitely not polarized. Finally, in
892: \Fignref{fig:T3GowdygeometricquantitiesHypVel} we show the
893: hyperbolic velocity, cf.\ \Sectionref{sec:gowdyphenom}. This is
894: computed from \Eqref{eq:hypvel} by first computing from our unknowns
895: an orthonormal frame as in \Sectionref{sec:comm_field_eqs} and then by
896: computing the relevant rescaled quantities. One sees that the velocity
897: nearly becomes constant in time which is a hint that we are not
898: approaching spiky features in the sense of
899: \Sectionref{sec:gowdyphenom}.
900:
901:
902: There are further reasons to believe that the upward pointing ``spiky''
903: features in
904: \Fignref{fig:T3GowdygeometricquantitiesKretsch} are just artifacts of
905: the gauge.
906: A simple argument is the
907: following. Polarized Gowdy spacetimes in areal gauge cannot develop
908: spikes which would be visible in the Kretschmann scalar due to the
909: results in \cite{Isenberg89}, meaning that the Kretschmann scalar
910: blows up uniformly when the singularity is approached.
911: However, the Kretschmann scalar of the polarized
912: solution corresponding to the polarized initial
913: data given at the beginning of this section is nearly
914: indistinguishable from
915: \Fignref{fig:T3GowdygeometricquantitiesKretsch}, although individual
916: variables are very different, see for instance\footnote{Note that the
917: jumps in this plot are
918: produced by our time adaption method described in
919: \Sectionref{sec:spatial_adaption}, since we plot the unrescaled
920: quantities here, and hence are not geometric.} \Figref{fig:T3Gowdye22}.
921: The relative deviation is of the order
922: $10^{-2}$ at the final time. In particular, the same upward pointing
923: features in the
924: Kretschmann scalar can be observed. This is a hint that the cause of
925: these features is the ``inhomogeneous'' approach to the
926: singularity in our gauge. However, we also cannot exclude the possibility
927: that at later times, even in this gauge, ``real'' spikes will be
928: visible.
929: In any case, it would be hard to
930: distinguish those ``real'' spikes from the effects that are caused by
931: the gauge.
932: \begin{figure}[tb]
933: \centering
934: \includegraphics[width=0.49\linewidth]
935: {plot_T3Sing_e22Pol_psfrag}
936: \caption{Evolution of non-rescaled $e\indices{_2^2}$ for a polarized and
937: non-polarized solution}
938: \label{fig:T3Gowdye22}
939: \end{figure}
940:
941:
942: I should note that when I stopped
943: the runs there were no principal numerical problems with the
944: solutions. Indeed, the evolutions could have continued closer to
945: the singularity.
946: The only limiting factor so far is the constraint growth. As
947: discussed before, I suspect that this is not
948: caused primarily by my discretization scheme but is rather a problem
949: either of the evolution equations at the continuum level or of the
950: gauge. We will discuss this problem again later.
951:
952: \subsection{Runs with \texorpdfstring{$\S$}{S3}-topology}
953: \label{sec:S3singularGowdy}
954:
955: In this section we report on similar investigations in the case of
956: $\S$-topology as before. Maybe one should note that these are the first
957: published attempts to study $\S$-Gowdy singularities numerically.
958: Recall from \Sectionref{sec:implementationOfS3Gowdy} that these are $2+1$ runs
959: in contrast to the $\T$-case and that the ideas to reduce to $1+1$
960: have not
961: been implemented yet. This means that we have higher practical
962: constraints on the spatial resolution now than in the
963: previous section. The
964: solutions constructed here also have to be seen in relation to those
965: in \Sectionref{sec:comparison} where we chose initial in the de-Sitter
966: stability region and hence obtained solutions which are both future
967: and past asymptotically de-Sitter. Here now, we want to leave the
968: stability region such
969: that the solutions become singular in the past. All runs in this
970: section have been done with the ``D2U Stag.'' method
971: (\Sectionref{sec:comparison}).
972:
973: Two sets of initial data as constructed in
974: \Sectionref{sec:solelectrconstraintscri} are considered: Gowdy data
975: with
976: \begin{enumerate}
977: \item ``small inhomogeneity''
978: \[a_3=1,\,a_3=0.7,\,(E_{11})_{0,0}=(E_{22})_{0,0}=0,\,C_2=10^{-4},\]
979: \item ``large inhomogeneity''
980: \[a_3=1,\,a_3=0.7,\,(E_{11})_{0,0}=(E_{22})_{0,0}=0,\,C_2=10^{-1}.\]
981: \end{enumerate}
982: As always so far, we assume that we are in
983: Levi-Civita conformal Gauß gauge.
984:
985: \begin{figure}[tb]
986: \centering
987: \includegraphics[width=0.49\linewidth]
988: {plot_S3Sing_AdaptPol_psfrag}
989: \caption{Behavior of $\normadapt$ for large and small
990: inhomogeneity}
991: \label{fig:S3Gowdyadapt}
992: \end{figure}
993: It turns out that both cases appear to become singular at
994: $t\approx 0.7$ because some variables blow up. The behavior of the
995: Kretschmann scalar and other quantities is discussed in a moment. As
996: expected, the
997: ``singular'' times are a bit different in the two
998: simulations. The runs
999: where done with the adaption mechanisms described in
1000: \Sectionref{sec:spatial_adaption};
1001: \Figref{fig:S3Gowdyadapt} shows the behaviors of
1002: $\normadapt$ in both cases. The
1003: $t$-axis has been stretched exponentially such that one can see the
1004: exponentially increasing dynamics close to the singularity in both
1005: cases. For the large inhomogeneity run, the
1006: adapted resolution, referred to as hShT, starts with $N_1=13$, $N_2=7$ and
1007: $h=2.5\cdot 10^{-4}$ and ends with $N_1=157$, $N_2=79$ and
1008: $h=3.90625\cdot 10^{-7}$. The resolutions lShT etc.\ are derived from
1009: that as
1010: above. Here $N_1$ is the number of collocation points in
1011: $\chi$-direction and $N_2$ in $\rho_1$-direction.
1012: For the small inhomogeneity case, the resolutions hShT starts with $N_1=13$,
1013: $N_2=7$ and $h=2.5\cdot 10^{-4}$ and ends with $N_1=469$, $N_2=235$ and
1014: $h=3.75\cdot 10^{-6}$. The reason that the resolution for the small
1015: inhomogeneity case ended up higher than for the large inhomogeneity
1016: case -- the other way around would have certainly made more sense -- was my
1017: unskilful choice of representative variable to compute
1018: $\normadapt$, namely, $E_{11}$ in both runs. From
1019: \Eqsref{eq:n2solelectrconstraint} we see that the initial data
1020: parameter $C_2$ controls the magnitude of the initial values of
1021: $E_{11}$. Although the initial value of $\normadapt$ was almost
1022: identical in both cases because of the definition of the norm, the
1023: consequence of this was that the time behavior of $\normadapt$
1024: was very different in both cases. I tried to compensate this by giving
1025: different threshold values for the adaption, see again
1026: \Fignref{fig:S3Gowdyadapt}, however, the undesired result was that
1027: the low inhomogeneity run was done with higher resolution than the
1028: high inhomogeneity run. The fact, as discussed below, that the error
1029: quantities in the two runs
1030: are almost of identical size suggests on the one hand, that the low
1031: inhomogeneity run did not really require so much resolution, but on the
1032: other hand, that the code is also not instable when the resolution is
1033: too high (at least so far). One could have worried about this issue, cf.\ the
1034: investigations related to
1035: \Figref{fig:linearized_instability}.
1036:
1037:
1038: \begin{figure}[tb]
1039: \centering
1040: \subfloat[Behavior of $\normelec$]{
1041: \includegraphics[width=0.49\linewidth]{plot_S3Sing_ConstrE_Large_psfrag}}%
1042: \subfloat[Behavior of $\normeinstein$]{
1043: \includegraphics[width=0.49\linewidth]
1044: {plot_S3Sing_EinsteinNorm_Large_psfrag}}\par
1045: \subfloat[Behavior of $\normkilling$]{%
1046: \label{fig:S3GowdyerrorKilling}
1047: \includegraphics[width=0.49\linewidth]{plot_S3Sing_YKilling_Large_psfrag}}
1048: \caption{Some error norms for the large inhomogeneity case}
1049: \label{fig:S3Gowdyerror}
1050: \end{figure}
1051: In \Figref{fig:S3Gowdyerror} we see the behavior of some error norms
1052: vs.\ time in the large inhomogeneity case. The behavior is
1053: analogous to the $\T$-case before. The errors are
1054: moderate until very close to the singularity; at the end of the run
1055: $\normelec$ is of the order $\sim 10^{-6}$ and $\sim 10^{-8}$ for
1056: $\normeinstein$ for the hShT run. Note however that
1057: here, in contrast to the $\T$-case before, the relevant components of
1058: the solutions are of the order $10^1$ at the final time (as we do not
1059: show here) and hence,
1060: the relative errors indeed grow close the singularity, but are still
1061: acceptable. The suspected
1062: reason for this is the limited spatial resolution since these are
1063: $2+1$-runs in contrast to $1+1$-runs before. Further, it is
1064: interesting that $\normkilling$ in \Fignref{fig:S3GowdyerrorKilling}
1065: is very stable until the errors in the solution start to grow
1066: more rapidly close to the singularity. We come back to this at the end
1067: of this section.
1068: As mentioned already
1069: above, the corresponding error
1070: quantities for the small inhomogeneity case behave very similar and thus
1071: are not plotted here.
1072:
1073: \begin{figure}[tb]
1074: \centering
1075: \subfloat[Orbit volume density]{%
1076: \label{fig:S3GowdygeometricquantitiesLargeOVol}
1077: \includegraphics[width=0.49\linewidth]{plot_S3Sing_OVol_Large_1D_psfrag}}%
1078: \subfloat[Kretschmann scalar]{%
1079: \label{fig:S3GowdygeometricquantitiesLargeKretsch}
1080: \includegraphics[width=0.49\linewidth]
1081: {plot_S3Sing_Kretsch_Large_1D_psfrag}}\par
1082: \subfloat[Scalar product of Killing vector fields]{%
1083: \label{fig:S3GowdygeometricquantitiesLargeScalarprKV}
1084: \includegraphics[width=0.49\linewidth]
1085: {plot_S3Sing_KVScalarpr_Large_1D_psfrag}}
1086: \caption{Spatial behavior of certain geometric quantities for large
1087: inhomogeneity}
1088: \label{fig:S3GowdygeometricquantitiesLarge}
1089: \end{figure}
1090: \begin{figure}[tb]
1091: \centering
1092: \subfloat[Orbit volume density]{%
1093: \label{fig:S3GowdygeometricquantitiesSmallOVol}
1094: \includegraphics[width=0.49\linewidth]{plot_S3Sing_OVol_Small_1D_psfrag}}%
1095: \subfloat[Kretschmann scalar]{%
1096: \label{fig:S3GowdygeometricquantitiesSmallKretsch}
1097: \includegraphics[width=0.49\linewidth]
1098: {plot_S3Sing_Kretsch_Small_1D_psfrag}}
1099: \caption{Spatial behavior of certain geometric quantities for small
1100: inhomogeneity}
1101: \label{fig:S3GowdygeometricquantitiesSmall}
1102: \end{figure}
1103: Next consider \Figref{fig:S3GowdygeometricquantitiesLarge} where we
1104: plot the spatial dependency of certain geometric
1105: quantities for five different times close to the singularity for
1106: the large inhomogeneity case. Corresponding plots for the small
1107: inhomogeneity case can be found in
1108: \Figref{fig:S3GowdygeometricquantitiesSmall}. Let us start with the
1109: rescaled orbit volume density\footnote{By ``rescaled'' we mean
1110: that the quantity is divided by $\sin^22\chi$ as explained in
1111: \Sectionref{sec:S3orbitvoletc}.}
1112: in
1113: \Fignref{fig:S3GowdygeometricquantitiesLargeOVol}. Note, as in the
1114: $\T$-case, that this
1115: quantity would be constant on each $t=const$ slice in areal gauge,
1116: however in our gauge it is not. In
1117: the plot we shift the values of the functions such that
1118: the most right point agrees for all curves
1119: so that we can study the deformations of the curves
1120: on the way to the ``singularity''. Similar to the $\T$-case, these curves
1121: deform such that points which are closer to the singularity
1122: move faster towards it which is caused by the Gauß like gauge. A
1123: similar behavior can be recognized for the small
1124: inhomogeneity case
1125: \Figref{fig:S3GowdygeometricquantitiesSmallOVol}. Now consider
1126: \Figref{fig:S3GowdygeometricquantitiesLargeKretsch} which shows the
1127: spatial dependence of the physical Kretschmann scalar. We see that
1128: with increasing time a localized feature develops as we observed also
1129: in the $\T$-case in
1130: \Fignref{fig:T3GowdygeometricquantitiesKretsch}. However, nothing
1131: like this is visible in the small inhomogeneity case
1132: \Fignref{fig:S3GowdygeometricquantitiesSmallKretsch}. In fact, the contrary
1133: seems to be the case there, namely the curves seem to become flatter
1134: with increasing time.
1135:
1136: What is happening here? When the
1137: inhomogeneity parameter $C_2$ of the initial data, which has the value
1138: $10^{-4}$ in the small inhomogeneity case and $10^{-1}$ in the large
1139: inhomogeneity case, is turned to zero, the corresponding solution is a
1140: $\lambda$-Taub-NUT spacetime with a Cauchy horizon in the past, and
1141: expressed in our variables and gauge,
1142: some quantities, in particular the trace of the $2$nd fundamental form,
1143: blow up there. This is so since the leaves approach a null
1144: surface. However, the Kretschmann scalar stays bounded.
1145: The behavior of the small inhomogeneity solution here shows very
1146: similar behavior indeed; for instance, the trace of the $2$nd fundamental
1147: form blows up (which we do not show here), but the Kretschmann scalar seems
1148: to stay bounded as indicated by
1149: \Figref{fig:S3GowdygeometricquantitiesSmallKretsch}. Now, we could
1150: speculate that the small inhomogeneity case also develops a Cauchy
1151: horizon in the past. In contrast, the large
1152: inhomogeneity case shows first signs that the Kretschmann
1153: scalar blows up, see
1154: \Figref{fig:S3GowdygeometricquantitiesLargeKretsch}. If this
1155: speculation was true then the Cauchy horizon of the
1156: $\lambda$-Taub-NUT spacetime given by zero inhomogeneity parameter
1157: $C_2$ would be stable under small inhomogeneous Gowdy perturbations of our
1158: type. Then, the small inhomogeneity case, provided the generators of
1159: the null-hypersurface are closed, should fit into a
1160: generalization of the family of solutions in \cite{moncrief84}, and it
1161: would be interesting to find
1162: out how the relation is. There is no simple a priori way of bringing
1163: the two pictures
1164: together, since Moncrief studies the problem as a singular initial value
1165: problem with the Cauchy horizon as initial hypersurface while we start
1166: from $\scrip$.
1167: In any case, all this is just speculation since so far our results are
1168: not conclusive. In fact it might
1169: turn out that other curvature quantities than the Kretschmann scalar,
1170: which we do not monitor currently, blow up in the small inhomogeneity
1171: run. On the other
1172: hand, because the inhomogeneity is so small it might also be the case that
1173: the Kretschmann scalar blows up at just a little later time. In any case, a
1174: systematic study
1175: of these issues is in order. Even if it turns out that the small
1176: inhomogeneity
1177: case does not
1178: correspond to a spacetime with Cauchy horizon or that, even more, the
1179: Cauchy horizon of the corresponding $\lambda$-Taub-NUT spacetime is
1180: not stable under these kind of perturbations at all, it is still interesting
1181: to study the transition from our inhomogeneous spacetime family with
1182: curvature singularities to a
1183: $\lambda$-Taub-NUT spacetime with Cauchy horizon. Further one should
1184: investigate other classes of perturbations. The perturbations
1185: considered so far are in the Gowdy class. According to the results
1186: listed in \Sectionref{sec:CHcosm} there must not be smooth Cauchy
1187: horizons in classes without symmetry. It would be interesting to study
1188: what happens when we systematically reduce the symmetry assumptions
1189: for our perturbations.
1190:
1191: There is another remarkable aspect of
1192: \Fignref{fig:S3GowdygeometricquantitiesLargeKretsch}. The maximum of
1193: the Kretschmann scalar at the latest considered times does not
1194: correspond to the place which is closest to the singularity in
1195: contrast to the $\T$-case. We could speculate if this is a spike
1196: in the sense of \Sectionref{sec:gowdyphenom} but further
1197: investigations of this are clearly necessary. Finally,
1198: \Fignref{fig:S3GowdygeometricquantitiesLargeScalarprKV}
1199: shows the scalar product of the Killing fields (similar for the
1200: small inhomogeneity) proving that the spacetime is not
1201: polarized.
1202:
1203: Unfortunately, considerations about the interesting non-linear stability
1204: issue of the class of $\S$-Gowdy spacetimes within the class of
1205: $\U$-symmetric spacetimes had a slightly lower priority in the
1206: presentation here. What we find numerically in the
1207: singular class studied in this section, as
1208: indicated in \Fignref{fig:S3GowdyerrorKilling} for the large
1209: inhomogeneity case, is consistent with what we found in the
1210: regular class in \Sectionref{sec:comparison}. Namely, the Gowdy symmetry of
1211: the numerical solutions is quite stable until the errors close to the
1212: singularity become dominant. A
1213: strong non-linear
1214: instability of the continuum equations would most likely have been
1215: visible in the
1216: numerical results. Hence, our findings indicate that such an
1217: instability is not present. Certainly, for reliable conclusions,
1218: further investigations are necessary.
1219:
1220: \section{Runs with the commutator field equations}
1221: \label{sec:RunsCosmFE}
1222:
1223: \subsection{Introduction}
1224: In the previous sections we presented some numerical calculations of
1225: singular $\lambda$-Gowdy spacetimes in the two cases of $\T$- and
1226: $\S$-topology using the conformal field equations in Levi-Civita
1227: conformal Gauß gauge. It turned out that this gauge is not
1228: well adapted to studies of the corresponding singularities because
1229: on the one hand the singularities are approached in a non-homogeneous way
1230: which makes interpretation of the results difficult; on the other hand
1231: the solution demands more and more resolution on time scales which
1232: become exponentially shorter when approaching the singularity.
1233: Since from previous experience we know that such
1234: problems do not
1235: occur so strongly
1236: with the commutator field
1237: equations in timelike area gauge (\Sectionref{sec:comm_field_eqs}),
1238: this
1239: motivates us to start numerical experiments.
1240: Numerical results with similar equations based on finite differencing
1241: methods can be
1242: found in \cite{Andersson03} in the
1243: case of a non-vanishing cosmological constant. Our first aim in this section
1244: is to get a feeling how well our pseudospectral approach can cope
1245: with the demands of this class of spacetimes under these ``better''
1246: gauge conditions. Another aim
1247: is to deepen the fundamental understanding of the Gowdy case with non-vanishing
1248: $\lambda$ by direct comparisons with the $\lambda=0$-case. Indeed, the
1249: investigations in the following are, to my
1250: knowledge, the only published numerical attempts to treat the Gowdy
1251: case with $\lambda\not=0$. However, due to time
1252: constraints in this thesis work, I only computed one numerical result
1253: regarding this issue which I present here,
1254: and otherwise just elaborate on my expectations. We start off by
1255: discussing suitable error analysis and error monitoring for this
1256: system.
1257:
1258: Unfortunately the
1259: formulation of the commutator field equations is restricted to the
1260: case of $\T$-topology and
1261: the modification to $\S$-topology is outstanding.
1262:
1263: Note that the solutions considered in this section are the only ones
1264: in this thesis which are not necessarily FAdS. This is so because we
1265: give data on standard Cauchy surfaces, and this yields no direct control about
1266: the evolution behavior. For $\lambda=0$, corresponding solutions cannot
1267: be FAdS anyway. In the following let us choose the time orientation
1268: such that the solutions collapse into the future, as determined by the
1269: timelike area gauge with positive $\mathcal N_0$ and increasing time $t$.
1270:
1271: \subsection{Error analysis and monitoring}
1272: \label{sec:erroranalcommruns}
1273: In \Chapterref{ch:numexperiments} we obtained some preliminary
1274: experience on the numerical behavior of the conformal field equations and fixed
1275: a few error quantities that we monitored in the
1276: discussion which followed. Since the situation is a bit different for
1277: the commutator
1278: field equations, this section is devoted to the discussion of some error
1279: quantities for these equations.
1280:
1281: Let us comment on how to judge the size of the errors involved in our
1282: numerical runs. First, this can be done as usual by convergence
1283: tests but, as explained before, their interpretation can be quite
1284: different for pseudospectral than for finite differencing codes due to
1285: the potentially strong influence of round-off errors in the first case.
1286: Second, one can monitor error quantities, in
1287: particular constraint violations.
1288: We have from \Eqref{eq:constrLambda}
1289: \[{\mathcal C}_{\Lambda}:=(E_1-2r)(\Omega_\Lambda)\]
1290: with $r$ obtained from \Eqref{eq:ralb} with $A\equiv 0$
1291: \[r = - 3\,(N_\times\,\Sigma_--N_-\,\Sigma_\times).\]
1292: Furthermore, we have from \Eqsref{eq:constrEresidual}
1293: \begin{align*}
1294: {\mathcal C}\indices{_2^2}&:=(E_1-\sqrt{3} N_\times-r)(E\indices{_2^2})\\
1295: {\mathcal C}\indices{_3^2}&:=(E_1+\sqrt{3} N_\times-r)(E\indices{_3^2})
1296: +2\sqrt{3} N_-E\indices{_2^2}\\
1297: {\mathcal C}\indices{_3^3}&:=(E_1+\sqrt{3} N_\times-r)(E\indices{_3^3}).
1298: \end{align*}
1299: The Gauß constraint \Eqref{eq:constrGauss} is solved identically since
1300: $\Sigma_+$ is obtained from it. However, the evolution equation for
1301: $\Sigma_+$ \Eqref{eq:evolSigmaPlus}
1302: \[3\,E_0(\Sigma_+)
1303: = -3\,(q+3\Sigma_+)\,(1-\Sigma_+)
1304: + 6\,(\Sigma_++\Sigma_-^{2}+\Sigma_\times^{2})- 3\,\Omega_\Lambda
1305: -E_1(r)\]
1306: might be violated numerically.
1307: How do we measure the violation of this equation at a given time step?
1308: There is no unique
1309: way to evaluate the partial derivatives in this equation
1310: numerically. The cleanest way, i.e.\ avoiding further errors which are just
1311: caused by the numerical evaluation of the partial derivatives, is to
1312: substitute the relation for $\Sigma_+$ from the Gauß constraint and
1313: the relation for $r$ above into this equation. Then, the equation
1314: involves time derivatives of other unknowns. With the same
1315: argument as above, it is natural to determine
1316: these values from the corresponding evolution equations whose
1317: violations at a given time step cannot be measured without introducing
1318: further distinct numerical methods. It turns out that all terms cancel and
1319: from this point of view the evolution equation for $\Sigma_+$ above is
1320: satisfied identically. This is no surprise since it is just a
1321: reformulation of the statement that the Gauß constraint propagates. In
1322: summary, we cannot measure the violation of the evolution equation of
1323: $\Sigma_+$ at a given time step without introducing further numerical
1324: methods to estimate the partial derivatives involved. In general, it is a
1325: principle that ``we cannot measure errors with methods that rely on
1326: the same errors''. The usage of other numerical methods to evaluate
1327: partial derivatives introduces further errors and it is hard to
1328: distinguish if the corresponding results really represent the true
1329: error quantity or rather the errors of these further numerical
1330: methods. For our purposes, we thus ignore the evolution equation of
1331: $\Sigma_+$.
1332:
1333: Now we comment on violations of the integrability condition
1334: \Eqref{eq:integrabbeta} for $\beta$ and define its violation as
1335: $\Phi_{int}$. After
1336: the same kind of manipulations as before we find
1337: \[\Phi_{int}=\frac{3}{2}{\mathcal C}_{\Lambda}.\]
1338: Hence, this error quantity is non-trivial, but it
1339: is explicitly determined by ${\mathcal C}_{\Lambda}$ and so it is
1340: sufficient to monitor the quantity ${\mathcal C}_{\Lambda}$.
1341:
1342: Further, we would like to check if the
1343: orbit area density
1344: $\mathcal A$ given by \Eqref{eq:orbitareadensity} is really constant
1345: on each time slice as required by the underlying gauge conditions. We
1346: define the following error quantity
1347: \[\mathcal C_A:=E_1(\mathcal A)
1348: =E_1(\beta^2 E\indices{_2^2}E\indices{_3^3}),\]
1349: and want to check if it is zero.
1350: The same kind of manipulations as above lead to
1351: \[\mathcal C_A=\mathcal A^{-1}(
1352: {\mathcal C}\indices{_2^2}+{\mathcal C}\indices{_3^3}).\]
1353: Hence, monitoring the quantities ${\mathcal C}\indices{_2^2}$ and
1354: ${\mathcal C}\indices{_3^3}$ is sufficient to estimate the error
1355: quantity $\mathcal C_A$.
1356:
1357: In summary, it is sufficient to monitor the following ``constraint
1358: violation'' quantity in our computations which is
1359: the sum of the $L^1$-norms of the quantities
1360: ${\mathcal C}_{\Lambda}$, ${\mathcal C}\indices{_2^2}$,
1361: ${\mathcal C}\indices{_3^2}$ and ${\mathcal C}\indices{_3^3}$.
1362:
1363:
1364: For the following runs, it turns out that no time adaption is
1365: needed. This can be expected
1366: since the gauge is chosen such that the singularity lies at
1367: $t\rightarrow\infty$ in an ``exponential manner''; see the discussion
1368: associated with the choice of lapse in \Eqref{eq:choicelapse}. A
1369: typical plot of the adaption norm is \Figref{fig:CosmBothadapt}, where
1370: one can see
1371: that the need for spatial resolutions indeed increases in time, but
1372: not on shorter and shorter time scales as in the runs before.
1373:
1374: \subsection{Numerical Results}
1375: \label{sec:commNumR}
1376: As initial data for the $\beta$-normalized quantities
1377: (\Sectionref{sec:comm_field_eqs}) we pick
1378: \[E\indices{_1^1}=-2, \quad\Sigma_-=-\frac 5{\sqrt 3}\cos x,
1379: \quad \Sigma_\times=0,\quad N_-=\frac 1{\sqrt 3}\sin x,\quad N_\times=0,\]
1380: which so far agrees with those in \cite{Andersson03} and which is in the
1381: same family of data that was already investigated in \cite{Berger93}. However,
1382: the residual choices are $\Omega_\Lambda=1$ (refer to as
1383: ``$\lambda>0$'') and
1384: $\Omega_{\Lambda}=0$ (refer to as ``$\lambda=0$'') which both satisfy
1385: the constraint \Eqref{eq:constrLambda}. Furthermore, the initial data
1386: for the residual decoupled part is
1387: \[E\indices{_2^2}=-2,\quad E\indices{_3^2}=2 \cos x,\quad
1388: E\indices{_3^3}=-2\]
1389: which is in agreement with the constraints
1390: \Eqsref{eq:constrEresidual}. Note that these data imply the initial
1391: hyperbolic velocity $v=5|\cos x|$ according to
1392: \Eqref{eq:hypvel}. Further recall that
1393: due to our specific choice of $\mathcal N_0=-1$, our
1394: time coordinate $t$ is related to the time coordinate $\tilde t$ in
1395: \cite{Andersson03} by $t=\tilde t/2$.
1396:
1397: \begin{figure}[tb]
1398: \begin{minipage}[t]{0.49\linewidth}
1399: \centering
1400: \includegraphics[width=\linewidth]
1401: {plot_T3Cosm_both_adapt_psfrag}
1402: \caption{Adapt norm for the $\lambda>0$ and $\lambda=0$ cases}
1403: \label{fig:CosmBothadapt}
1404: \end{minipage}
1405: \begin{minipage}[t]{0.49\linewidth}
1406: \centering
1407: \includegraphics[width=\linewidth]
1408: {plot_T3Cosm_both_ConstrE_psfrag}
1409: \caption{Behavior of the constraint violation for the $\lambda>0$
1410: and $\lambda=0$ cases}
1411: \label{fig:Cosmconstraint}
1412: \end{minipage}
1413: \end{figure}
1414: In \Figref{fig:CosmBothadapt} we see the behavior of $\normadapt$. As
1415: indicated before due to the choice of gauge, there is no need to
1416: stretch the time axis. Certainly
1417: this subserves the quality of the numerical calculations since no
1418: ``artificial'' time adaption is necessary.
1419: In the following we show only results obtained with one resolution and
1420: in particular no convergence plots; the code is convergent similarly as
1421: we observed before.
1422: The time resolution is
1423: $h=10^{-4}$ in all runs shown here and fixed; the spatial resolution starts with
1424: $N=511$ (number of collocation points in $x$-direction) in both runs
1425: and ends with $N=2133$ in the $\lambda>0$ case and with $3779$ in the
1426: $\lambda=0$-case.
1427: In both cases,
1428: the final time $t=1.8999$ corresponds
1429: to $\tilde t=3.7998$ (time coordinate in \cite{Andersson03}). This is
1430: not very close to the singularity; note, however that there was no
1431: principal obstacle to let the runs continue.
1432:
1433: In \Figref{fig:Cosmconstraint} one can see the behavior of the
1434: constraint violation as introduced in
1435: \Sectionref{sec:erroranalcommruns}. We see that the constraints are
1436: surprisingly well behaved,
1437: namely they decay approximately exponentially in time. Note that these jumps in
1438: \Fignref{fig:Cosmconstraint} are produced by the spatial
1439: adaption. Namely, the
1440: individual $L^1$-norms of the violations
1441: are not exactly equal before and after an interpolation step. In any
1442: case, this is a big
1443: difference to the results which we obtained with the conformal field
1444: equations.
1445:
1446: \begin{figure}[tb]
1447: \centering
1448: \subfloat[Behavior of $\Sigma_\times$]{
1449: \includegraphics[width=0.49\linewidth]{plot_T3Cosm_nonvanish_u6_1D_psfrag}}%
1450: \subfloat[Behavior of $N_-$]{
1451: \includegraphics[width=0.49\linewidth]
1452: {plot_T3Cosm_nonvanish_u7_1D_psfrag}}\par
1453: \subfloat[Behavior of $\beta$]{%
1454: \label{fig:CosmNonvanishGeomBeta}
1455: \includegraphics[width=0.49\linewidth]{plot_T3Cosm_nonvanish_u9_1D_psfrag}}%
1456: \subfloat[Behavior of hyperbolic velocity]{
1457: \includegraphics[width=0.49\linewidth]{plot_T3Cosm_nonvanish_HypVel_1D_psfrag}}
1458: \caption{$\lambda>0$ case: geometric quantities for $5$
1459: different times}
1460: \label{fig:CosmNonvanishGeom}
1461: \end{figure}
1462: \begin{figure}[tb]
1463: \centering
1464: \subfloat[Behavior of $\Sigma_\times$]{
1465: \includegraphics[width=0.49\linewidth]{plot_T3Cosm_vanish_u6_1D_psfrag}}%
1466: \subfloat[Behavior of $N_-$]{
1467: \includegraphics[width=0.49\linewidth]
1468: {plot_T3Cosm_vanish_u7_1D_psfrag}}\par
1469: \subfloat[Behavior of hyperbolic velocity]{
1470: \includegraphics[width=0.49\linewidth]{plot_T3Cosm_vanish_HypVel_1D_psfrag}}
1471: \caption{$\lambda=0$ case: geometric quantities for $5$
1472: different times}
1473: \label{fig:CosmvanishGeom}
1474: \end{figure}
1475: Now consider \Figref{fig:CosmNonvanishGeom} and
1476: \Figref{fig:CosmvanishGeom} where we plot the spatial dependence of
1477: some geometric quantities for a few time steps for the $\lambda>0$-
1478: and the $\lambda=0$-case respectively. From these plots we can follow
1479: how the solution develops localized features from smooth data. In both
1480: cases a particularly sharp feature develops at $x=1$, particularly
1481: visible in the $\Sigma_\times$ plots, but which then decays
1482: again. Comparing the $\lambda=0$ plots to Fig.~10 of
1483: \cite{Andersson03},
1484: which shows\footnote{But note that the
1485: authors of \cite{Andersson03} use Hubble-normalized quantities so that
1486: the plots cannot be compared directly.}
1487: $\Sigma_\times$ at $t=5$ ($\tilde t=10$), we see
1488: that there is indeed no ``final'' spike at $x=1$. Those can be
1489: observed more on the left and
1490: symmetrically on the right of $x=1$ and we can speculate that our
1491: plots show the early stages of such in both cases
1492: $\lambda>0$ and $\lambda=0$. Consider also
1493: \Fignref{fig:CosmNonvanishGeomBeta}
1494: which has little peaks at those expected positions. But why does the
1495: feature at $x=1$ decay in both cases? It is a ``high velocity spike''
1496: \cite{Garfinkle03} since its initial hyperbolic velocity is bigger
1497: than $2$. The phenomenology of these features can be described roughly
1498: as follows. The evolution
1499: equations drive them to lower velocity while some of them (as in our case)
1500: decay completely. In \cite{Garfinkle03}, $\lambda=0$ is assumed but in
1501: this special case here we see that the presence
1502: of $\lambda$ does not change this behavior very much. However, the
1503: cosmological constant seems to lead to faster decay. An
1504: intuitive argument for this is that the repulsive forces of the
1505: cosmological constant blow up the localized features. Consider
1506: \Fignref{fig:CosmBothadapt} again. Here we can follow the formation and
1507: decay of this intermediate spiky feature in frequency space. We see
1508: that the fine structure is
1509: built in short phases, shorter for $\lambda>0$, with relaxation phases
1510: in between. This is consistent with the investigations of
1511: \cite{Garfinkle03} where the conclusion was drawn that the high
1512: velocity spikes decay by bouncing from one velocity regime to the next
1513: lower, while in between there are phases of relatively weak
1514: dynamics. In any case, we can see that plotting $\normadapt$ is not only
1515: interesting to study, how the code itself behaves, but also how the
1516: solution develops. This is so since building fine structure requires higher
1517: resolutions and this is represented well in Fourier space. Indeed, one
1518: can view this as an advantage
1519: of pseudospectral methods compared to other methods because the role
1520: of frequency space, in which part of the relevant information about fine
1521: structure can be read off directly, plays a fundamental role for the
1522: method itself.
1523:
1524: These intermediate
1525: ``spikes'' demand quite high spatial resolutions already after a short time.
1526: But after that, when these spikes have decayed again, the resolution
1527: is not needed anymore at least for some
1528: time. The current implementation of my spatial
1529: adaption method cannot cope with
1530: this situation very well; in particular it is not able to reduce the
1531: resolution when it is not needed
1532: anymore. However, it should be straight forward to modify the method
1533: to make this possible. In any case, these intermediate spikes make my
1534: runs quite slow already after short time. Although the runs are not
1535: yet impractically slow I stopped them so that I was not able to
1536: study the final distribution of the spikes. Another reason for
1537: stopping the
1538: runs is that the current implementation of the
1539: code produces much output data, of the order of $10$ GByte, and
1540: the hard disk that I was using ran out of free space.
1541:
1542: \begin{figure}[tb]
1543: \centering
1544: \subfloat[Behavior of $N_-$]{
1545: \includegraphics[width=0.49\linewidth]{plot_T3Cosm_vanish_u7_lowvel_1D_psfrag}}%
1546: \subfloat[Behavior of Hyperbolic velocity]{
1547: \includegraphics[width=0.49\linewidth]
1548: {plot_T3Cosm_vanish_HypVel_lowvel_1D_psfrag}}
1549: \caption{Low velocity run}
1550: \label{fig:CosmvanishGeomLowVel}
1551: \end{figure}
1552: However,
1553: since the intermediate spikes are caused by the high
1554: velocity of my choice of initial data sets, I also
1555: computed the solution of another choice of
1556: initial data with lower initial velocity.
1557: This choice is
1558: the same as above but $\Omega_{\Lambda}=0$ and
1559: $\Sigma_-=-\frac 2{\sqrt 3}\cos x$, so that the initial hyperbolic velocity is
1560: $v=2|\cos x|$. In the corresponding \Figref{fig:CosmvanishGeomLowVel}
1561: we see that there is actually a spike at $x=1$ (and also at $x=0$)
1562: which does not decay and we can see its development exemplary for
1563: $N_-$. Note, that the time scales are different for these data than
1564: before and we stopped the code at $t=4.5999$. The
1565: final spatial resolution was $N=2133$.
1566:
1567:
1568: \subsection{Expectations regarding \texorpdfstring{$\Omega_{\Lambda}$}
1569: {Omega(Lambda)}}
1570: \label{sec:ExpectT3GowdyLambda}
1571: The numerical studies before cannot be considered as systematic
1572: investigations of the influence of the cosmological
1573: constant in solutions with Gowdy symmetry. The only aspect we could
1574: see is that maybe $\lambda>0$ supports the decay of high velocity
1575: spikes. But what about low velocity spikes? Could there be an extreme
1576: value of $\Omega_\Lambda$ such that all spikes decay before the singularity?
1577:
1578: The only analytical results known so far for $\T$-Gowdy with
1579: $\lambda>0$ are due to
1580: Clausen and Isenberg \cite{Clausen07} who prove that the maximal
1581: Cauchy development of any smooth Gowdy initial data on a
1582: standard Cauchy surface is globally
1583: covered by areal coordinates where the orbit area lies in the interval
1584: $]c,\infty[$ for some undetermined constant $c\ge 0$. Furthermore,
1585: by means of Fuchsian methods as in
1586: \cite{Kichenassamy97} they obtain that one can construct solutions which are
1587: asymptotically velocity dominated in the analytic case. However, one
1588: does not obtain
1589: control over the full set of free functions.
1590:
1591: We can say a little more than this, although many of the following
1592: arguments are heuristic. \Eqref{eq:qgauge} for $q$ tells
1593: us that
1594: \[q>\frac 12-\frac 32\Omega_\Lambda,\]
1595: when we exclude the fixpoint solution
1596: $\Sigma_-=N_\times=\Sigma_\times=N_-=0$ (de-Sitter spacetime).
1597: Then, with $\mathcal N_0=-1$, \Eqref{eq:OmegaLambdaDot} implies that
1598: \[\dot\Omega_\Lambda<3(\Omega_\Lambda-1)\,\Omega_\Lambda.\]
1599: Consider the following initial value problem
1600: \[\dot y(t)=3(y(t)-1)y(t),\quad y(0)=\eta.\]
1601: A unique solution exists which has the explicit form
1602: \[y(t)=\frac 1{1-C e^{3t}}\]
1603: with $C=1-1/\eta$. The solution exists for all $t\ge 0$ if and only if
1604: $\eta\le 1$. If $\eta<1$, $y(t)=O(e^{-3t})$ for
1605: $t\rightarrow\infty$. Now since, $\Omega_{\Lambda}$ is a subsolution
1606: of this problem it follows from the standard theory of ODEs that for
1607: $\Omega_{\Lambda}(0)<1$ we have $\Omega_{\Lambda}(t)=O(e^{-3t})$ for
1608: $t\rightarrow\infty$. If $\Omega_{\Lambda}(0)=1$ we can only deduce
1609: that $\Omega_{\Lambda}(t)<1$ for all $t$. Now for
1610: $\Omega_{\Lambda}\ll 1$ for late $t$ we can expect from \Eqref{eq:E11dot} that
1611: $E\indices{_1^1}=O(e^{-2t})$ for
1612: $t\rightarrow\infty$ and hence $\Omega_{\Lambda}$ decays exponentially
1613: faster than $E\indices{_1^1}$. Since roughly speaking, the decay of
1614: $E\indices{_1^1}$ is responsible for bringing the solution into the
1615: asymptotically velocity dominated regime and also for spiky features
1616: we can expect the same phenomenology as in the $\lambda=0$-case, at
1617: least if $\Omega_{\Lambda}(0)<1$.
1618:
1619: What about very large $\Omega_{\Lambda}$? If it is initially large
1620: compared to the other unknowns then
1621: $q\approx -\frac 32\Omega_{\Lambda}$.
1622: Using this in the evolution equation \Eqref{eq:OmegaLambdaDot}
1623: implies
1624: \[\dot\Omega_\Lambda\approx 3 \Omega_\Lambda\Omega_\Lambda,\]
1625: which has a solution that is unbounded after finite time. Let us assume
1626: that the other unknowns are so small initially such that in particular
1627: the quadratic terms can be neglected. Then all of the evolution
1628: equations are of the form
1629: $\dot u/u\approx 3\Omega_\Lambda$ for a generic unknown $u$. Comparing
1630: this with the equation
1631: for $\Omega_\Lambda$ this could mean that all these
1632: quantities increase with the same strength such that the approximation
1633: $q\approx -\frac 32\Omega_{\Lambda}$ is still valid for later
1634: times. Then the whole
1635: solution could blow up after finite time. However, it is not clear in
1636: particular how the non-linear terms in the evolution equations behave
1637: in this situation. If
1638: it turns out that the other unknowns grow faster initially than
1639: $\Omega_{\Lambda}$, then $q$ would become more and more positive
1640: and the grow of $\Omega_{\Lambda}$ would be damped such that possibly the
1641: solution stays finite for all times. In any case, what I want to say
1642: by means of
1643: this heuristic discussion is that is not easy, without usage of more difficult
1644: arguments, to exclude a blow up after finite time if the
1645: $\Omega_{\Lambda}$ is
1646: high enough initially. Such a blow up would imply a drastic change in the
1647: dynamics for $t\rightarrow\infty$ compared to the well known
1648: $\lambda=0$ case. This possibility is maybe related to the outstanding
1649: issues in the
1650: theorems by Clausen et al.\ explained above. Can the constant $c$
1651: always be chosen to be
1652: zero? As we said, for $\lambda$ ``small enough'', this can be
1653: done. But if not, what kind of singular behavior is there?
1654: Note, that if the solution is, say, past
1655: asymptotically de-Sitter (past $\leftrightarrow$ decreasing the
1656: $t$-coordinate), then
1657: all timelike
1658: geodesics are future incomplete in this case, at least when the scalar
1659: curvature of $\scrip$ is negative, due to the singularity theorems of
1660: \Sectionref{sec:singularitytheoremGA}.
1661: But do they run into curvature
1662: singularities or Cauchy horizons? In the first case: are the solution
1663: in some sense asymptotically velocity dominated? Recall, that Clausen
1664: et al.\ were not able to prove that \textit{all} solutions are
1665: asymptotically velocity dominated, not even
1666: in the analytic
1667: case.
1668:
1669:
1670:
1671:
1672:
1673: