gr-qc0210022/1.tex
1: \documentclass{article} 
2:   
3: \usepackage{amssymb,eufrak,epsfig,latexsym} 
4: \parindent0mm 
5: \setlength{\parskip}{1ex} 
6: 
7: %\input /afs/aei-potsdam.mpg.de/u/jurke/latex/kwg/prom/Doktorarbeit/Latex/STYLE-FI/tcilatex
8: \input tcilatex 
9: \newenvironment{beweis}{{\bfseries\upshape\selectfont Proof:}} 
10:  {\ifmmode\rlap{\hskip\displaywidth\rule{2mm}{2mm}}\else\hfill $\rule{2mm}{2mm}$\fi} 
11: \def\ABB#1#2#3{\vskip6mm\def\Test{#2}\ifx\Test\empty\centerline{%
12:  \epsfig{figure=#1}}\else\centerline{\epsfig{figure=#1, width=#2}}\fi%
13:  \def\Test{#3}\ifx\Test\empty\else\vskip0mm\global\advance\c@Abbildung %
14:  by 1\def\@currentlabel{(\thechapter-\theAbbildung)}%
15:  \centerline{\mbox{\rmfamily\mdseries\upshape\fontsize{9pt}{13pt}%
16:  \selectfont \begin{tabular}{rl} Abbildung (\thechapter-\theAbbildung):%
17:  \kern -3mm & #3\end{tabular}}}\fi\vskip5mm}
18: 
19: \begin{document} 
20: 
21: \title{On future asymptotics of polarized Gowdy  
22: $\mathbb T^3$-models} 
23: \author{Thomas Jurke\\
24: {\em Max-Planck-Institut f\"ur Gravitationsphysik}\\ 
25: {\em Am M\"uhlenberg 1}\\ 
26: {\em 14476 Golm}\footnote{current address: 
27: Weierstra{\ss}-Institut 
28: f\"ur Angewandte Analysis und 
29: Stochastik, Mohrenstra{\ss}e 39, 10117 Berlin}}
30: \maketitle 
31: 
32: \begin{abstract} 
33: Gowdy's model of cosmological spacetimes is a much investigated 
34: subject in classical and quantum gravity. Depending on spatial 
35: topology re\-collapsing as well as expanding models are known. 
36: Several analytic tools were used in order to clarify singular 
37: behaviour in this class of spacetimes. Here we investigate the 
38: structure of a certain subclass, the polarized Gowdy models with 
39: spatial $\mathbb T^3$-topology, in the large. 
40: The a\-sym\-pto\-tics for general solutions of the dynamical equation for one 
41: of the gravitational degrees of freedom plays a key role 
42: while the asymptotic behaviour of the remaining metric function 
43: is a result of solving the Hamiltonian constraint equation. 
44: Using both we are able to prove (future) geodesic 
45: completeness in all spacetimes of this type. 
46: \end{abstract} 
47: 
48:  
49: \section{Introduction} 
50: %\begin{eqnarray} 
51: %\frac{\partial^2}{\partial t^2}W\kern1mm+\kern1mm
52: %\frac1t\frac{\partial}{\partial t}W\kern1mm-\kern1mm
53: %\frac{\partial^2}{\partial x^2}W&=&\sinh W\cosh W
54: %\left(\left(\frac{\partial Q}{\partial t}\right)^2-
55: %\left(\frac{\partial Q}{\partial x}\right)^2\right)
56: %	\label{GowdyW}\\ 
57: %\frac{\partial^2}{\partial t^2}Q\kern1mm+\kern1mm
58: %\frac1t\frac{\partial}{\partial t}Q\kern1mm-\kern1mm
59: %\frac{\partial^2}{\partial x^2}Q&=&-2\coth W
60: %\left(\frac{\partial Q}{\partial t}\frac{\partial W}{\partial t}-
61: %\frac{\partial Q}{\partial x}\frac{\partial W}{\partial x}\right)
62: %        \label{GowdyQ}\\ 
63: %\frac{\partial^2}{\partial t^2}a\kern1mm+\kern1mm
64: %\frac1t\frac{\partial}{\partial t}a\kern1mm-\kern1mm
65: %\frac{\partial^2}{\partial x^2}a&=&\frac12
66: %\left(\left(\frac{\partial W}{\partial x}\right)^2+\sinh^2 W
67: %\left(\frac{\partial Q}{\partial x}\right)^2\right)
68: %        \label{Gowdya}
69: %\end{eqnarray}
70: %\begin{eqnarray}
71: %\frac{\partial}{\partial t}a&=&
72: %-\frac1{4t}+\frac{t}4\left[
73: %\left(\frac{\partial W}{\partial t}\right)^2+
74: %\left(\frac{\partial W}{\partial x}\right)^2+
75: %\right.\nonumber\\
76: %&&\left.
77: %\sinh^2 W\left(
78: %\left(\frac{\partial Q}{\partial t}\right)^2+
79: %\left(\frac{\partial Q}{\partial x}\right)^2
80: %\right)
81: %\right]
82: %	\label{Gowdyat}\\ 
83: %\frac{\partial}{\partial x}a&=&\frac{t}2\left(
84: %\frac{\partial W}{\partial t}
85: %\frac{\partial W}{\partial x}+
86: %\sinh^2 W\frac{\partial Q}{\partial t}
87: %\frac{\partial Q}{\partial x}
88: %\right)		
89: %	\label{Gowdyax} 
90: %\end{eqnarray} 
91: The cosmological spacetime model by Gowdy (1974) has experienced a great
92: deal of attention over the last years. These models are especially 
93: interesting because they allow rigorous analytical investigations in
94: inhomogeneous spacetimes. Various spatial topologies are allowed but~--~as
95: is mostly done~--~we will concentrate on vacuum spacetimes with $\mathbb
96: R_+\times\mathbb T^3$-topology. A basic analytical tool for this class of
97: spacetimes was provided by Moncrief~(1981) who proved existence of global
98: smooth solutions for the corresponding Einstein equations. Investigations 
99: concerning the singular asymptotic behaviour were subjects of interest
100: almost exclusively. Now, we have a quite clear picture what happens
101: in the polarized as well as the full Gowdy $\mathbb T^3$-model (see
102: Isenberg, Moncrief~(1990), Rendall~(2000), Ringstr\"om~(2002) and references 
103: therein). On the other hand, due to Chru\'sciel, Isenberg and Moncrief~(1990)
104: there exist some statements about the asymptotic behaviour for large times in 
105: the case of polarized spacetimes. In this paper we prove such theorems, 
106: describing the long time asymptotics for general polarized Gowdy models
107: and using this results to prove future completeness of any causal geodesic.
108: 
109: 
110: \section{The polarized model}  
111: By definition a smooth, orientable spacetime is a cosmological model of Gowdy
112: type if it is a maximally extended, globally hyperbolic solution of the
113: vacuum Einstein equations having compact Cauchy surfaces and a 
114: pseudo-Riemannian metric, invariant under an effective $U(1)\times U(1)$ 
115: group action on these Cauchy surface. This action is generated by two
116: spacelike Killing vector fields which are assumed to have vanishing twist
117: constants. A Gowdy model is called polarized if the defining Killing vector
118: fields are everywhere mutually orthogonal. By virtue of this definition
119: we may describe a polarized Gowdy metric by the line element
120: \begin{eqnarray}
121: {\rm d}s^2&=&{\rm e}^{2a(t,x)}\left(-{\rm d}t^2+{\rm d}x^2\right)\kern1mm+
122: \kern1mmt\kern.5mm\left({\rm e}^{W(t,x)}{\rm d}y^2+{\rm e}^{-W(t,x)}
123: {\rm d}z^2\right).
124: \end{eqnarray}
125: The corresponding vacuum Einstein equations are
126: \begin{eqnarray} 
127: 0&=&\frac{\partial^2}{\partial t^2}W\kern1mm+\kern1mm 
128: \frac1t\frac{\partial}{\partial t}W\kern1mm-\kern1mm 
129: \frac{\partial^2}{\partial x^2}W\label{Darboux}\\[2mm] 
130: 0&=&\frac{\partial^2}{\partial t^2}a\kern1mm-\kern1mm 
131: \frac{\partial^2}{\partial x^2}a 
132: \kern1mm-\kern1mm\frac1{4t^2}\kern1mm+\kern1mm\frac14 
133: \left[\left(\frac\partial{\partial t}W\right)^2\kern1mm-\kern1mm 
134: \left(\frac\partial{\partial x}W\right)^2 
135: \right]\label{Wellena}\\[2mm] 
136: 0&=& 
137: \frac2t\frac{\partial}{\partial x}a\kern1mm-\kern1mm 
138: \frac{\partial}{\partial t}W\kern1mm\frac{\partial}{\partial x}W 
139: \label{Impuls1}\\[2mm] 
140: 0&=& 
141: \frac2t\frac{\partial}{\partial t}a\kern1mm+\kern1mm 
142: \frac1{2t^2}\kern1mm-\kern1mm\frac12\left[\left( 
143: \frac{\partial}{\partial t}W\right)^2\kern1mm+\kern1mm 
144: \left(\frac{\partial}{\partial x}W\right)^2\right]. 
145: \label{Hamilton1} 
146: \end{eqnarray} 
147: As in the general non-polarized case equation~(\ref{Wellena}) is always 
148: satisfied as 
149: a consequence of the others. Thus, we have essentially to deal with a single 
150: linear equation of second order while  
151: function~$a$ is the integral of the constraint  
152: equation~(\ref{Hamilton1}). 
153:   
154: \section{On the asymptotics for solutions of the  
155: Euler-Poisson-Darboux equation} 
156:  
157: Now, we start by investigating equation~(\ref{Darboux}). As in 
158: literature we will sometimes call this equation 
159: ``Euler-Poisson-Darboux equation''.  
160: Following Moncrief~(1981) it will be sufficient to have only 
161: functions $W=W(t,x)$ of a certain regularity in mind
162: which are defined for all 
163: $(t,x)\in\mathbb R_+\times\mathbb T^1$. 
164:   
165: As a first result we get asymptotic homogenity for all 
166: solutions of~(\ref{Darboux}). Namely, we will show 
167: that for large~$t$ the mean value, 
168: \begin{eqnarray} 
169: \bar W(t) &\mathop{=}\limits_{\rm Df}& \frac1{2\pi} 
170: \intop\limits_0^{2\pi} W(t, x)\kern1mm {\rm d} x, 
171: \label{W_mittel} 
172: \end{eqnarray}  
173: will dominate while the $x$-dependent terms show a  
174: $t^{-\frac12}$ fall-off behaviour. 
175:  
176: \begin{theorem} 
177: Let $W\in\mathcal C^2(\mathbb R_+\times\mathbb T^1;\mathbb R)$  
178: be a solution of Euler-Poisson-Darboux equation~(\ref{Darboux}) 
179: and $\bar W$ as decleared in~(\ref{W_mittel}). Then $\bar W$ 
180: solves~(\ref{Darboux}), too. 
181: \end{theorem} 
182:  
183: \begin{beweis} 
184: For every $t\in\mathbb R_+$ function $W$ as well as its first and 
185: second derivative are continuous, hence 
186: $\mathbb T^1$-integrable and $\bar W$ is well-defined.  
187: Since~$W$ is sufficiently smooth $\bar W$ depends on its 
188: parameter~$t\in\mathbb R_+$ smoothly, hence
189: $\bar W\in\mathcal C^2(\mathbb R_+;\mathbb R)$ and 
190: \begin{eqnarray*}  
191: \frac{\rm d^2}{{\rm d}t^2}\bar W(t)&=&\frac1{2\pi}  
192: \frac{\rm d}{{\rm d}t}\int_0^{2\pi}\frac\partial{\partial t}  
193: W(t,x)\kern1mm{\rm d}x\kern2mm=\kern2mm  
194: \frac1{2\pi}\int_0^{2\pi}\frac{\partial^2}{\partial t^2}  
195: W(t,x)\kern1mm{\rm d}x.  
196: \end{eqnarray*} 
197: Now it is easy to integrate~(\ref{Darboux})  
198: \begin{eqnarray*} 
199: \frac1{2\pi} 
200: \intop\limits_0^{2\pi} 
201: \frac{\partial^2}{\partial t^2}W \kern1mm {\rm d} x 
202: \kern2mm+\kern2mm 
203: \frac1{2\pi t} 
204: \intop\limits_0^{2\pi} 
205: \frac{\partial }{\partial t}W\kern1mm {\rm d} x &=& 
206: \frac1{2\pi} 
207: \intop\limits_0^{2\pi} 
208: \frac{\partial ^2}{\partial x^2}W\kern1mm {\rm d} x\\[2mm] 
209: %\end{eqnarray*} 
210: %\begin{eqnarray*} 
211: \frac{\rm d^2}{{\rm d} t^2}\bar W \kern2mm+\kern2mm 
212: \frac1t\frac{\rm d }{{\rm d} t}\bar W&=& 
213: \frac1{2\pi}\left. 
214: \frac{\partial W}{\partial x}\right|^{2\pi}_0 
215: \end{eqnarray*} 
216: and since the derivative of smooth periodic functions have same 
217: periods we get 
218: \begin{eqnarray*} 
219: \frac{\partial^2}{\partial t^2}\bar W \kern2mm+\kern2mm 
220: \frac1t\frac{\partial }{\partial t}\bar W&=&0\kern2mm 
221: \mathop{=}\limits^! \kern2mm 
222: \frac{\partial^2}{\partial x^2}\bar W. 
223: \end{eqnarray*} 
224: \end{beweis} 
225:  
226: The general solution of~(\ref{Darboux}) which does not depend on $x$
227: is 
228: \begin{eqnarray} 
229: \bar W(t)&=&\gamma\kern1mm+\kern1mm 
230: \beta\cdot\ln t, \label{Whom} 
231: \end{eqnarray} 
232: where $\beta$ and $\gamma$ may be fixed by some initial values 
233: of~$\bar W$ and~${\bar W}_t$ at $t_0>0$. Due to the 
234: linearity of~(\ref{Darboux}) $\psi$ in
235: \begin{eqnarray*} 
236: t^{-\frac12}\psi(t,x)&\mathop{=} 
237: \limits_{\rm Df}& W(t,x)-\bar W(t) 
238: \end{eqnarray*} 
239: is unique and solves (\ref{Darboux}), too. Obviously this 
240: newly defined function~$\psi$,   
241: $\psi\in\mathcal C^2(\mathbb R_+\times\mathbb T^1;\mathbb R)$,  
242: has zero mean value for every~$t\in\mathbb R_+$  
243: \begin{eqnarray} 
244: \bar\psi(t)&\mathop{=}\limits_{\rm Df}& 
245: \frac1{2\pi}\intop\limits_0^{2\pi}\psi(t,x)\kern1mm{\rm d}x 
246: \kern2mm=\kern2mm0\label{mittellos} 
247: \end{eqnarray} 
248: and satisfies 
249: \begin{eqnarray} 
250: \frac{\partial^2}{\partial t^2}\psi \kern1mm-\kern1mm 
251: \frac{\partial ^2}{\partial x^2}\psi&=& 
252: -\frac1{4t^2}\psi. 
253: \label{Wellenpsi} 
254: \end{eqnarray} 
255:   
256: Now, in order to prove that any solution of this equation has 
257: to be bounded we need the following estimate as prerequisite.  
258: Here we have formulated the lemma in dependence of a 
259: parameters~$t\in\mathbb R_+$, since this is the form we will need. 
260: \begin{lemma}  
261: Let \label{Hilfssatz2}  
262: $f\colon{\mathbb R}_+\times {\mathbb T}^1\to {\mathbb R}$ be 
263: continuously differentiable and  
264: \begin{eqnarray*} 
265: \frac1{2\pi}\int_0^{2\pi} f(t,x)\kern1mm {\rm d} x&=&0 
266: \end{eqnarray*} 
267: for every $t$. Then we have for each $(t,x)\in\mathbb R_+ 
268: \times\mathbb T^1:$
269: \begin{eqnarray*}  
270: |f(t,x)|^2 &\le& 2\pi\intop\limits_0^{2\pi}\left|  
271: \frac\partial{\partial x} f(t,x)  
272: \right|^2\kern1mm {\rm d} x\kern2mm\equiv\kern2mm 2\pi\kern1mm 
273: \left|\left| f_x(t,.)\right|\right|_{L^2}^2  
274: \end{eqnarray*}  
275: \end{lemma}  
276:   
277: \begin{beweis}  
278: The mean value theorem of integral calculus for continuous 
279: functions gives us to any~$t$ some 
280: $x_0 \in {\mathbb T}^1$ with $f(t, x_0)=0$. Furthermore   
281: $|x-x_0|\le 2\pi$ is true for each $x\in\mathbb T^1$. Using 
282: Schwartz' inequality we get 
283: \begin{eqnarray*}  
284: |f(t,x)| &=&   
285: \left|\intop\limits_{\phantom{.}x_0}^{x}  
286: \frac\partial{\partial\xi} f(t,\xi)  
287: \kern1mm {\rm d} \xi \right|			  
288: \kern2mm\le\kern2mm \intop\limits_0^{2\pi}\left|  
289: \frac\partial{\partial x} f(t,x)  
290: \right|\kern1mm {\rm d} x			\\  
291: &\le& \left(\intop\limits_0^{2\pi}1^2  
292: \kern1mm {\rm d} x\right)^{\frac12}\cdot  
293: \left(  
294: \intop\limits_0^{2\pi}\left|  
295: \frac\partial{\partial x} f(t,x)  
296: \right|^2\kern1mm {\rm d} x  
297: \right)^{\frac12}  
298: \end{eqnarray*}  
299: that is 
300: \begin{eqnarray*}  
301: |f(t,x)|^2 &\le& 2\pi\intop\limits_0^{2\pi}\left|  
302: \frac\partial{\partial x} f(t,x)  
303: \right|^2\kern1mm {\rm d} x.  
304: \end{eqnarray*}  
305: \end{beweis}  
306:  
307: Now we are able to show boundedness of~$\psi$, 
308: uniformly on every strictly positive intervall. The key element 
309: in the proof is the investigation of a certain energy 
310: functional corresponding to~(\ref{Wellenpsi}). 
311:   
312: \begin{theorem}  
313: To every function~$\psi\in\mathcal C^2(\mathbb R_+ 
314: \times\mathbb T^1;\mathbb R)$, satisfying both~(\ref{mittellos})  
315: and~(\ref{Wellenpsi}), there is a positive constant~$C_{t_0}$ 
316: depending only on~$t_0>0$, such that for every~$t\ge t_0$ and  
317: all $x\in\mathbb T^1$ the inequality 
318: \label{Satz4} 
319: \begin{eqnarray}  
320: |\psi(t, x)|<C_{t_0} \label{Schaetz} 
321: \end{eqnarray}  
322: holds.
323: \end{theorem} 
324:   
325: \begin{remark} 
326: One verifies that equation (\ref{Wellenpsi}) has unbounded 
327: solutions, too. But none of them satisfies~(\ref{mittellos}) 
328: which underlines the importance of this assumption.
329: \end{remark}  
330:  
331: \begin{beweis} 
332: Let us define the energy 
333: \begin{eqnarray*} 
334: \varepsilon(t)&\mathop{=}\limits_{\rm Df}&\frac12  
335: \int_0^{2\pi}\left({\psi_t}^2+{\psi_x}^2\right){\rm d}x. 
336: \end{eqnarray*}   
337: Our smoothness assumptions in~$\psi$ guarantees 
338: differentiability of~$\varepsilon$ in $\mathbb R_+$.  
339: Integrating by parts and using equation~(\ref{Wellenpsi}) yields 
340: \begin{eqnarray*}  
341: \frac{\rm d}{{\rm d} t}\varepsilon&=& 
342: -\frac1{4 t^2}\int \limits_0^{2\pi} 
343: \psi_t \psi\kern1mm{\rm d}x. 
344: \end{eqnarray*}  
345: Since $\frac12({\psi_t}^2+\psi^2)+\psi_t\psi\ge0$ we have 
346: \begin{eqnarray*} 
347: \frac{\rm d}{{\rm d} t}\varepsilon  
348: &\le& 
349: \frac1{4 t^2}\cdot\frac12\int \limits_0^{2\pi} 
350: \left({\psi_t}^2+\psi^2\right)\kern1mm{\rm d}x.  
351: \end{eqnarray*}  
352: Now, lemma~\ref{Hilfssatz2} permits an estimate of the left 
353: hand side of the inequality in terms of $\varepsilon$ itself: 
354: \begin{eqnarray*} 
355: \frac{\rm d}{{\rm d} t}\varepsilon&\le&\frac{\pi^2}{t^2}\varepsilon  
356: \end{eqnarray*}  
357: Since $t\ge t_0>0$ we find  
358: \begin{eqnarray*}  
359: \varepsilon (t)&\le&  
360: \varepsilon (t_0)\kern1mm+\kern1mm  
361: \pi^2\intop\limits_{t_0}^t s^{-2}  
362: \varepsilon(s)\kern1mm {\rm d}s 
363: \end{eqnarray*}  
364: and Gronwall's inequality gives the uniform bound  
365: \begin{eqnarray*}  
366: \varepsilon (t)&\le&  
367: \varepsilon (t_0)\kern1mm+\kern1mm  
368: \varepsilon (t_0)\left({\rm e}^{\frac{\pi^2}{t_0}  
369: -\frac{\pi^2}{t\phantom{_0}}}  
370: -1\right)  
371: \kern2mm<\kern2mm  
372: \varepsilon (t_0)\kern1mm  
373: {\rm e}^{\frac{\pi^2}{t_0}}.  
374: \end{eqnarray*}  
375: Hence, the energy of an arbitrary solution is for every 
376: $t>t_0$ bounded and by definition non-negative.  
377: Now, up to a factor we can control the  
378: $L^2(\mathbb T^1)$-norm 
379: of~$\psi_x$ for all $t\ge t_0$ by the square root of the same 
380: constant and since~$\psi$ satisfies all suppositions of  
381: lemma~\ref{Hilfssatz2} a short calculation yield finally 
382: \begin{eqnarray*} 
383: |\psi(t,x)|&<&\sqrt{4\pi\kern.5mm\varepsilon(t_0)}\kern1mm  
384: {\rm e}^{\frac{\pi^2}{2t_0}}\kern2mm\equiv\kern2mm C_{t_0} 
385: \end{eqnarray*}  
386: for all $t\ge t_0>0$ and all $x\in\mathbb T^1$, hence $C_{t_0}$  
387: fulfills all needs. 
388: \end{beweis}  
389:  
390: We have shown that every spatially periodic solution of the 
391: Euler-Poisson-Darboux equation can be cast into the form 
392: \begin{eqnarray*} 
393: W(t,x)&=&\gamma\kern1mm+\kern1mm\beta\cdot\ln t 
394: \kern1mm+\kern1mmt^{-\frac12}\psi(t,x) 
395: \end{eqnarray*} 
396: where $\psi$ is bounded in $t\in[t_0,\infty)$. On the other 
397: hand for applications in the following sections we need deeper 
398: insights into the asymptotic behaviour of~$\psi$. 
399: Using the boundedness of~$\psi$ we can show that solutions 
400: to 
401: \begin{eqnarray*} 
402: \frac{\partial^2}{\partial t^2}\psi\kern1mm-\kern1mm 
403: \frac{\partial^2}{\partial x^2}\psi&=& 
404: -\frac1{4t^2}\psi 
405: \end{eqnarray*} 
406: behave like solutions of the free wave equation in two 
407: dimensions with a remainder falling off in time. 
408:   
409: \begin{theorem}  
410: Let $\psi\in\mathcal C^2(\mathbb R_+\times\mathbb T^1;\mathbb R)$  
411: be a bounded solution of \label{Satz5} 
412: \begin{eqnarray}  
413: \frac{\partial^2}{\partial t^2}\psi\kern1mm-\kern1mm 
414: \frac{\partial^2}{\partial x^2}\psi&=& 
415: -\frac1{4t^2}\psi,\label{Psi2} 
416: \end{eqnarray}  
417: with the side condition 
418: \begin{eqnarray*} 
419: \frac1{2\pi}\int_0^{2\pi}\psi(t,x)\kern1mm{\rm d}x&=&0 
420: \end{eqnarray*} 
421: for every $t\in\mathbb R_+$. Then there exist uniquely 
422: defined functions 
423: $\nu\in\mathcal C^2(\mathbb R_+\times\mathbb T^1;\mathbb R)$ and  
424: $\omega\in\mathcal C^2(\mathbb R_+\times\mathbb T^1;\mathbb R)$ 
425: as well as a constant~$C$ depending only on the choise of $t_0$  
426: such that  
427: \begin{eqnarray*}  
428: \psi(t,x)&=&\nu(t,x)\kern1mm+\kern1mm\omega(t,x)\\[2mm]  
429: \frac{\partial^2}{\partial t^2}\nu&=& 
430: \frac{\partial^2}{\partial x^2}\nu \\[2mm] 
431: |\omega(t,x)|&\le& C\cdot t^{-1}\\[2mm]  
432: |\omega_t (t,x)|&\le& C\cdot t^{-1}  
433: \end{eqnarray*} 
434: for all $t\in[t_0,\infty)$, $t_0>0$ and all $x\in\mathbb T^1$.  
435: \end{theorem} 
436:  
437: \begin{beweis}  
438: Let 
439: \begin{eqnarray*} 
440: u&\mathop{=}\limits_{\rm Df}&t\kern1mm+\kern1mm x\\[2mm] 
441: v&\mathop{=}\limits_{\rm Df}&t\kern1mm-\kern1mm x. 
442: \end{eqnarray*} 
443: In this coordinates~(\ref{Psi2}) reads 
444: \begin{eqnarray} 
445: \frac{\partial^2}{\partial u\kern1mm\partial v}\psi&=& 
446: -\frac14\frac{\psi}{(u+v)^2}.\label{Psi3} 
447: \end{eqnarray} 
448: As a consequence of this transformation some parallelogram,  
449: say~$\overline{1234}$ 
450: \ABB{x-t.eps}{10cm}  
451: {%Integrationsgebiet in\\ \phantom{eine kleine Verschiebung}  
452: %charakteristischen Koordinaten  
453: }  
454: \mbox{}\vskip-16mm  
455: \begin{eqnarray*}  
456: &&\kern45mm  
457: \begin{array}{ll}  
458: \left(t^{(1)}\equiv t_0,\right.&\left.x^{(1)}\equiv x_0=u_0-t_0  
459: \right)\\[2mm]  
460: \left(t^{(2)} = t_0,\right.&\left.x^{(2)}=x_0+2\pi  
461: \right)\\[2mm]  
462: \left(t^{(3)}= t_0+U,\right.&\left.x^{(3)}=x_0-U  
463: \right)\\[2mm]  
464: \left(t^{(4)}= t_0+U,\right.&\left.x^{(4)}=x_0+2\pi-U  
465: \right)  
466: \end{array}  
467: \end{eqnarray*}  
468: \vskip5mm  
469:  
470: has new vertices as follows: 
471: \ABB{u-v.eps}{10cm}  
472: {%Integrationsgebiet in\\ \phantom{eine kleine Verschiebung}  
473: %charakteristischen Koordinaten  
474: }  
475: \mbox{}\vskip-16mm  
476: \begin{eqnarray*}  
477: &&\kern45mm  
478: \begin{array}{ll}  
479: \left(u^{(1)}\equiv u_0,\right.&\left.v^{(1)}\equiv v_0=2t_0-u_0  
480: \right)\\[2mm]  
481: \left(u^{(2)} = u_0+2\pi,\right.&\left.v^{(2)}=v_0-2\pi  
482: \right)\\[2mm]  
483: \left(u^{(3)}= u_0,\right.&\left.v^{(3)}=v_0+2U  
484: \right)\\[2mm]  
485: \left(u^{(4)}= u_0+2\pi,\right.&\left.v^{(4)}=v_0-2\pi+2U  
486: \right)  
487: \end{array}  
488: \end{eqnarray*}  
489: \vskip7mm  
490:  
491: The periodicity in $x$-space is the identity 
492: \begin{eqnarray} 
493: (u,v)&\equiv&(u+2\pi n,v-2\pi n)\label{Periode} 
494: \end{eqnarray} 
495: for all $n\in\mathbb Z$ and the over-all condition 
496: $t>0$ converts to $u+v>0$. Both are generally assumed in what 
497: follows.~--~The estimate 
498: \begin{eqnarray*} 
499: |\psi_u(u,v_2)\kern1mm-\kern1mm 
500: \psi_u(u,v_1)|&\le&\intop_{v_1}^{v_2}\frac{|\psi|\kern1mm{\rm d}v} 
501: {(u+v)^2}\kern2mm\mathop{\le}\limits^{(\ref{Schaetz})}\kern2mm 
502: C_{t_0}\left(\frac1{u+v_1}\kern1mm-\kern1mm\frac1{u+v_2}\right) 
503: \end{eqnarray*} 
504: yields boundedness of $|\psi_u(u,v)|$ for any $u$ while 
505: $v\to\infty$. But the same inequality also proves with 
506:  $v_2=v_1+\Delta$, $\Delta>0$, the existence of a finite 
507: limit $F^\prime(u)$ for any $u$, hence 
508: \begin{eqnarray} 
509: F^\prime(u)&\mathop{=}\limits_{\rm Df}&   
510: \psi_u(u,2t_0-u)+\lim_{U\to\infty} 
511: \intop_{2t_0-u}^{2t_0-u+2U} \psi_{uv}(u,v) 
512: \kern1mm {\rm d}v \label{FStr} 
513: \end{eqnarray} 
514: is well-defined. Periodicity of $F^\prime(u)$ is due  
515: to~(\ref{Periode}) along with 
516: \begin{eqnarray*}  
517: F^\prime(u+2\pi)&=&   
518: \psi_u(u+2\pi,2t_0-u-2\pi)-\frac14  
519: \intop_{2t_0-u-2\pi}^\infty \frac{\psi(u+2\pi,v)}  
520: {(u+2\pi+v)^2}  
521: \kern1mm {\rm d}v\\[2mm] 
522: &=&   
523: \psi_u(u,2t_0-u)-\frac14  
524: \intop_{2t_0-u}^\infty \frac{\psi(u+2\pi,\tilde v-2\pi)}  
525: {(u+\tilde v)^2}  
526: \kern1mm {\rm d}\tilde v\\  
527: &=& F^\prime(u). 
528: \end{eqnarray*} 
529: Another property of $F^\prime(u)$ is its zero mean value. 
530: To show this we transform the defining integral back to 
531: $t$-$x$-coordinates according to the scetches from the beginning 
532: of this proof. Let $(u_0,v_0)$, that is $(t_0,x_0)$, 
533: $t_0>0$, an arbitrary point in characteristic or space-time 
534: coordinates, respectively. Then 
535: \begin{eqnarray}  
536: \intop_{u_0}^{u_0+2\pi} F^\prime(u)\kern1mm {\rm d}u  
537: &=& -\intop_{u_0}^{u_0+2\pi}   
538: \lim\limits_{U\to\infty}\intop\limits_{2t_0-u}^{2t_0-u+2U}  
539: \frac {\psi(u,v)}{4(u+v)^2}  
540: \kern1mm {\rm d}v\kern1mm {\rm d}u	\\  
541: &=&\lim\limits_{U\to\infty}\intop_{t_0}^{t_0+U}\frac1{2t^2}   
542: \intop\limits_{x_0+t_0-t}^{x_0+t_0-t+2\pi}  
543: \psi(t,x)  
544: \kern1mm {\rm d}x\kern1mm {\rm d}t\kern2mm=\kern2mm0, 
545: \label{F-Periode}  
546: \end{eqnarray} 
547: since the term under the integral is zero for every~$t$.  
548: A direct consequence is the periodicity of primitives. Let 
549: \begin{eqnarray*} 
550: \tilde F(u)&\mathop{=}\limits_{\rm Df}&\int F^\prime(u)\kern1mm 
551: {\rm d}u 
552: \end{eqnarray*} 
553: such an (arbitrary but fixed) primitive. Then the function 
554: \begin{eqnarray*} 
555: F(u)&\mathop{=}\limits_{\rm Df}&\tilde F(u) 
556: \kern1mm-\kern1mm\frac1{2\pi}\intop_{u_0}^{u_0+2\pi}  
557: \tilde F(u)\kern1mm{\rm d}u 
558: \end{eqnarray*} 
559: is defined uniquely and especially 
560: \begin{eqnarray} 
561: \intop_{u_0}^{u_0+2\pi} F(u)\kern1mm 
562: {\rm d}u&=&0 \label{FS-Periode} 
563: \end{eqnarray} 
564: is always satisfied. 
565: Now, starting with the inequality 
566: \begin{eqnarray*} 
567: |\psi_v(u_2,v)\kern1mm-\kern1mm 
568: \psi_v(u_1,v)|&\le&\intop_{u_1}^{u_2}\frac{|\psi|\kern1mm{\rm d}u} 
569: {(u+v)^2}\kern2mm\le\kern2mm 
570: C_{t_0}\left(\frac1{u_1+v}\kern1mm-\kern1mm\frac1{u_2+v}\right) 
571: \end{eqnarray*} 
572: we construct a likewise uniquely defined function 
573: $G(v)$, where all steps follow as above 
574: endowing $G$ with corresponding properties. Let us finally 
575: introduce the functions: 
576: \begin{eqnarray} 
577: \nu(u,v)&\mathop{=}\limits_{\rm Df}& 
578: F(u)\kern1mm+\kern1mm G(v)\label{Nue}\\[2mm] 
579: \omega(u,v)&\mathop{=}\limits_{\rm Df}& 
580: \psi(u,v)\kern1mm-\kern1mm \nu(u,v) 
581: \end{eqnarray} 
582: Obviously $\nu$ and $\omega$ are uniquely defined for any fixed 
583: $\psi$. $\nu$ solves the free wave equation $\nu_{uv}=0$ 
584: and regarding the remainder~$\omega$ we estimate for every $u$, $v$ 
585: subject to $u+v>0$ 
586: \begin{eqnarray}  
587: |\omega_u(u,v)|&=&|\psi_u(u,v)-\nu_u(u,v)|		 
588: \nonumber\\ [2mm] 
589: &=&\left|\psi_u(u,v) - \psi_u(u,2t_0-u) -  
590: \intop_{2t_0-u}^\infty\psi_{uv}(u,v)\kern1mm{\rm d}v\right|  
591: \nonumber\\[2mm]  
592: &=&\left|-\intop^{2t_0-u}_v  
593: \psi_{u\tilde v}(u,\tilde v)\kern1mm  
594: {\rm d}\tilde v\kern1mm-\kern1mm  
595: \intop_{2t_0-u}^\infty\psi_{uv}(u,v)\kern1mm{\rm d}v\right|	 
596: \nonumber\\ [2mm] 
597: &=&  
598: \left|\frac14\intop_v^\infty  
599: \frac{\psi(u,\tilde v)}{(u+\tilde v)^2}\kern1mm{\rm d}  
600: \tilde v\right|\kern2mm\le\kern2mm\frac{C_{t_0}}{4(u+v)},		  
601: \label{Itom} 
602: \end{eqnarray} 
603: where the estimate follows from~(\ref{Schaetz}). In exactly the same 
604: way one proves 
605: \begin{eqnarray*}  
606: |\omega_v(u,v)|&\le&\frac{C_{t_0}}{4(u+v)}.  
607: \end{eqnarray*}  
608: Now, since
609: \begin{eqnarray}  
610: |\omega_t(t,x)|&=& 
611: |\omega_u(u,v)+\omega_v(u,v)| 
612: \kern2mm\le\kern2mm\frac {C_{t_0}}{2(u+v)}\kern2mm=\kern2mm 
613: \frac{C_{t_0}}{4t}\label{Itom2}\\[2mm]  
614: |\omega_x(t,x)|&=& 
615: |\omega_u(u,v)-\omega_v(u,v)| 
616: \kern2mm\le\kern2mm\frac {C_{t_0}}{2(u+v)}\kern2mm=\kern2mm 
617: \frac{C_{t_0}}{4t}\label{Itom1} 
618: \end{eqnarray}  
619: for all $t\in[t_0,\infty)$, $x\in\mathbb T^1$ we have 
620: \begin{eqnarray*} 
621: ||\omega_x(t,.)||_{L^2}^2&=& 
622: \intop_0^{2\pi} 
623: {\omega_x}^2\kern1mm{\rm d}x\kern2mm\le\kern2mm 
624: \frac{\pi {C_{t_0}}^2}{8t^2}. 
625: \end{eqnarray*} 
626: Corollary~\ref{om-Eigen} below proves the properties 
627: of~$\omega$ 
628: \begin{eqnarray*} 
629: \omega(t,x+2\pi)&=&\omega(t,x)	\\ 
630: \intop_0^{2\pi}\omega(t,x)\kern1mm{\rm d}x&=&0 
631: \end{eqnarray*} 
632: for every $(t,x)\in\mathbb R_+\times\mathbb T^1$. Hence all 
633: assumptions of lemma~\ref{Hilfssatz2} are satisfied which 
634: ensures the existence of some constant 
635: $C\mathop{=}\limits_{\rm Df}\frac\pi2 C_{t_0}>0$ with 
636: \begin{eqnarray} 
637: |\omega(t,x)|&\le& \frac Ct\label{A6} 
638: \end{eqnarray} 
639: for all $t\in[t_0,\infty)$, $t_0>0$ and all $x\in\mathbb T^1$. 
640: \end{beweis} 
641:  
642: The existence and uniqueness in the definition of function $\nu$ 
643: allow us to introduce the following term. 
644:  
645: \begin{definition}  
646: The function $\nu$ is called the  $\psi$-associated solution 
647: of the free wave equation, if~$\psi\in\mathcal C^2 
648: (\mathbb R_+\times\mathbb T^1;\mathbb R)$ solves 
649: \begin{eqnarray*} 
650: \frac{\partial^2}{\partial t^2}\psi\kern1mm-\kern1mm 
651: \frac{\partial^2}{\partial x^2}\psi&=& 
652: -\frac1{4 t^2}\psi 
653: \end{eqnarray*} 
654: with the side condition 
655: \begin{eqnarray*} 
656: \intop_0^{2\pi}\psi(t,x)\kern1mm{\rm d}x&=&0 
657: \end{eqnarray*} 
658: for every $t\in[t_0,\infty)$ and $\nu\in\mathcal C^2 
659: (\mathbb R_+\times\mathbb T^1;\mathbb R)$ is defined as in 
660: equation~(\ref{Nue}). 
661: \end{definition} 
662:  
663: Here follows a series of corollaries establishing properties 
664: of~$\nu$ as well as~$\omega$ which were used in the previous 
665: proof or are of frequent use in the sections below. 
666:  
667: \begin{corollary} 
668: Let $\nu$ be the $\psi$-associated solution of the free 
669: wave equation and $t_0>0$. Then there exists a positive constant 
670: $C$, such that for all $t\in[t_0,\infty)$ and $x\in\mathbb T^1$ 
671: the functions
672: $\nu$, $\nu_t$ and $\nu_x$ are bounded:\label{nu-Eigen} 
673: \begin{eqnarray} 
674: \mathop{\mathop{\rm max}\limits_{t\ge t_0>0}}\limits_{x\in\mathbb T^1} 
675: \kern1mm 
676: \left\{|\nu(t,x)|,|\nu_t(t,x)|,|\nu_x(t,x)|\right\}&<&C  \label{nuE1} 
677: \end{eqnarray} 
678: Further, for all $n\in\mathbb Z$ are: 
679: \begin{eqnarray} 
680: \nu(t,x+2\pi n)&=&\nu(t,x) 
681: \kern2mm=\kern2mm\nu(t+2\pi n,x)  \label{nuE2}	\\[2mm] 
682: \nu_t(t+2\pi n,x)&=&\nu_t(t,x)  \label{nuE4}	\\[2mm] 
683: \nu_x(t+2\pi n,x)&=&\nu_x(t,x)  \label{nuE5}	\\[2mm] 
684: \intop_0^{2\pi}\nu(t,x)\kern1mm{\rm d}x&=&0 \label{nuE3} 
685: \end{eqnarray} 
686: \end{corollary} 
687:  
688: \begin{beweis} 
689:  
690: (\ref{nuE1})\kern3mm 
691: By definition it is $\nu(u,v)=F(u)+G(v)$. $F$ is periodic with vanishing 
692: mean value. W.l.o.g.\ is $F(u_0)=0$, 
693: $\psi_u(u,v_\infty)\kern1mm\mathop{=}\limits_{\rm Df}\kern1mm 
694: \mathop{\rm lim}\limits_{v\to\infty}\psi_u(u,v)$. Then 
695: \begin{eqnarray*} 
696: F(u)&=&\intop_{u_0}^{u}F^\prime(\tilde u)\kern1mm{\rm d}\tilde u 
697: \kern2mm=\kern2mm 
698: \intop_{u_0}^{u}\psi_{\tilde u}(\tilde u,v_\infty) 
699: \kern1mm{\rm d}\tilde u 
700: \kern2mm=\kern2mm\psi(u,v_\infty)\kern1mm-\kern1mm\psi(u_0,v_\infty). 
701: \end{eqnarray*} 
702: Theorem~\ref{Satz4} shows boundedness of $\psi$  
703: (uniformly in the considered domain). 
704: The same is true for $G$. Finally, boundedness of $|\nu_t(t,x)|$ 
705: and $|\nu_x(t,x)|$ is a consequence of  
706: $|F^\prime(u)|$ and $|G^\prime(v)|$ boundedness established 
707: during the proof of theorem~\ref{Satz5}. 
708:  
709: (\ref{nuE2})\kern3mm 
710: With~(\ref{FS-Periode}) we have 
711: \begin{eqnarray*} 
712: \nu(t+2\pi n,x)&=&F(t+2\pi n+x)\kern1mm+\kern1mm G(t+2\pi n-x)\\ 
713: &=&F(t+x)\kern1mm+\kern1mmG(t-x) 
714: \kern2mm=\kern2mm\nu(t,x) 
715: \end{eqnarray*} 
716: and analogously 
717: \begin{eqnarray*} 
718: \nu(t,x+2\pi n)&=&F(t+x+2\pi n)\kern1mm+\kern1mm G(t-x-2\pi n)\\ 
719: &=&\nu(t,x). 
720: \end{eqnarray*} 
721:  
722: (\ref{nuE4})\kern3mm 
723: Is a direct consequence of (\ref{nuE2}). 
724:  
725: (\ref{nuE5})\kern3mm  
726: Follows directly from~(\ref{nuE2}). 
727:  
728: (\ref{nuE3})\kern3mm  
729: Property~(\ref{FS-Periode}) and correspondingly for~$G$ yield: 
730: \begin{eqnarray*} 
731: \intop_0^{2\pi} 
732: \nu(t,x)\kern1mm{\rm d}x&=& 
733: \intop_t^{t+2\pi} 
734: \nu(u,2t-u)\kern1mm{\rm d}u\\ 
735: &=& 
736: \intop_t^{t+2\pi}\left( 
737: F(u)\kern1mm+\kern1mm G(2t-u)\right) 
738: \kern1mm{\rm d}u\\ 
739: &=& 
740: \intop_t^{t+2\pi} 
741: F(u)\kern1mm{\rm d}u 
742: \kern1mm+\kern1mm 
743: \intop_{t-2\pi}^{t} G(v) 
744: \kern1mm{\rm d}v\kern2mm=\kern2mm0 
745: \end{eqnarray*} 
746: \end{beweis} 
747:  
748: \begin{corollary} 
749: Let $\nu$ be the $\psi$-associated solution of the free wave 
750: equation. Then the following are equivalent:\label{nu-Aequ} 
751: \begin{eqnarray} 
752: \psi(t,x)&\equiv&0    \label{nuA1}	\\[2mm] 
753: F(t+x)&\equiv&0 \kern5mm\text{and}\kern5mm 
754: G(t-x)\kern2mm\equiv\kern2mm0   \label{nuA2}	\\[2mm] 
755: \nu(t,x)&\equiv&0    \label{nuA3} 
756: \end{eqnarray} 
757: \end{corollary} 
758:  
759: \begin{beweis} 
760:  
761: (\ref{nuA1}) $\Rightarrow$ (\ref{nuA2}):\kern3mm  
762: $F$ is constant due to~(\ref{FStr}), 
763: from~(\ref{FS-Periode}) follows the statement with respect to~$F$;  
764: analogously for $G$. 
765:  
766: (\ref{nuA2}) $\Leftrightarrow$ (\ref{nuA3}):\kern3mm  
767: ($\Rightarrow$) is due  
768: to~(\ref{Nue}) and ($\Leftarrow$)~due to~(\ref{Nue}) 
769: and~(\ref{FS-Periode}). 
770:  
771: (\ref{nuA3}) $\Rightarrow$ (\ref{nuA1}):\kern3mm 
772: It is $\psi(t,x)=\omega(t,x)$. The estimate~(\ref{Itom1}) 
773: reads here in a first step 
774: \begin{eqnarray*} 
775: |\psi_x(t,x)|_1&\le&\frac{C_{t_0}}{4t} 
776: \end{eqnarray*} 
777: and according to lemma~\ref{Hilfssatz2} 
778: \begin{eqnarray*} 
779: |\psi(t,x)|_1&\le&\frac{\pi C_{t_0}}{2t}. 
780: \end{eqnarray*} 
781: Now, this improved estimate can be used iteratively in~(\ref{Itom}), 
782: leading to an $n$-th order estimate 
783: \begin{eqnarray*} 
784: |\psi(t,x)|_n&\le&\left(\frac\pi 2\right)^n \frac{C_{t_0}}{n!} 
785: t^{-n} 
786: \end{eqnarray*} 
787: for every $x\in\mathbb T^1$ and every $t\in[t_0,\infty)$. 
788: Due to~(\ref{Itom2}) the same estimate is valid for $\psi_t$.  
789: Consequently exists to any 
790: $\varepsilon>0$ an $n_0$, such that 
791: \begin{eqnarray*} 
792: |\psi(t_0,x)|_n&<&\varepsilon\\ 
793: |\psi_t(t_0,x)|_n&<&\varepsilon 
794: \end{eqnarray*} 
795: for all $x\in\mathbb T^1$ and all $n\ge n_0$, that is 
796: $\psi(t_0,x)\kern1mm=\kern1mm\psi_t(t_0,x)\kern1mm=\kern1mm0$. 
797: Due to the uniqueness of its solutions 
798: $\psi$ is the trivial solution of~(\ref{Psi2}). 
799: \end{beweis} 
800:  
801: \begin{corollary} 
802: Let $\nu$ be the $\psi$-associated solution of the free wave 
803: equation, $\omega\mathop{=}\limits_{\rm Df}\psi-\nu$  
804: and $t_0>0$. Then there exists a positive constant 
805: $C$, such that for all $t\in[t_0,\infty)$, $x\in\mathbb T^1$ 
806: and $n\in\mathbb Z$:\label{om-Eigen} 
807: \begin{eqnarray} 
808: |\omega(t,x)|&<&C  \label{omE1}	\\[2mm] 
809: \omega(t,x+2\pi n)&=&\omega(t,x)  \label{omE2}	\\[2mm] 
810: \intop_0^{2\pi}\omega(t,x)\kern1mm{\rm d}x 
811: &=&0  \label{omE3} 
812: \end{eqnarray} 
813: \end{corollary} 
814:  
815: \begin{beweis} 
816: All properties are direct consequences from the definition  
817: of~$\omega$, the corresponding properties of~$\psi$ along 
818: with corollary~\ref{nu-Eigen} above. 
819: \end{beweis} 
820:  
821: \begin{corollary} 
822: Let $\nu$ be the $\psi$-associated solution of the free wave 
823: equation and $\omega\mathop{=}\limits_{\rm Df}\psi-\nu$.  
824: Then the following is equivalent:\label{om-Aequ} 
825: \begin{eqnarray} 
826: \psi(t,x)&\equiv&0   \label{omA1}	\\[2mm] 
827: \omega(t,x)&\equiv&0 \label{omA2} 
828: \end{eqnarray} 
829: \end{corollary} 
830:  
831: \begin{beweis} 
832:  
833: (\ref{omA1}) $\Rightarrow$ (\ref{omA2}):\kern3mm  
834: With corollary~\ref{nu-Aequ} is $\nu\equiv 0$. 
835:  
836: (\ref{omA2}) $\Rightarrow$ (\ref{omA1}):\kern3mm  
837: It is $\psi\equiv\nu$. Since 
838: \begin{eqnarray*} 
839: \frac{\partial^2}{\partial t^2}\psi\kern1mm-\kern1mm 
840: \frac{\partial^2}{\partial x^2}\psi&=&-\frac1{4t^2}\psi 
841: \kern3mm\mathop{\equiv}\limits^{!}\kern3mm0 
842: \end{eqnarray*} 
843: the statement follows. 
844: \end{beweis} 
845:  
846: Now we summarize the most important features in the 
847: asymptotic behaviour of the metric function~$W$. 
848: Here we use as remainder  
849: $\varkappa\mathop{=}\limits_{\rm Df}t^{-\frac12}\omega$. 
850:   
851: \begin{corollary}\label{Folgerung6}%  
852: Let $W\in\mathcal C^2(\mathbb R_+\times \mathbb T^1)$  
853: be a real-valued solution of the Euler-Poisson-Darboux 
854: equation  
855: \begin{eqnarray*}  
856: \frac{\partial^2}{\partial t^2}W(t,x)\kern1mm+\kern1mm  
857: \frac1t  
858: \frac{\partial}{\partial t}W(t,x)  
859: \kern1mm-\kern1mm  
860: \frac{\partial^2}{\partial x^2}W(t,x)&=&0.  
861: \end{eqnarray*}  
862: Then there exists uniquely defined, bounded, real-valued 
863: functions $\nu\in\mathcal C^2(\mathbb R_+\times\mathbb T^1)$  
864: and $\varkappa\in\mathcal C^2(\mathbb R_+\times\mathbb T^1)$,  
865: constants $\beta$, $\gamma$, and a positive constant $C_{t_0}$ 
866: such that 
867: \begin{eqnarray}  
868: W(t,x)&=&\gamma\kern1mm+\kern1mm\beta\cdot\ln t  
869: \kern1mm+\kern1mm t^{-\frac12}\nu(t,x)\kern1mm+\kern1mm  
870: \varkappa(t,x)\label{A1}\\  
871: \frac{\partial^2}{\partial t^2}\nu(t,x)&=&  
872: \frac{\partial^2}{\partial x^2}\nu(t,x)\label{A2}  
873: \end{eqnarray}  
874: and  
875: \begin{eqnarray}  
876: |\varkappa(t,x)|&\le&C_{t_0}\cdot t^{-\frac32}\label{A3}\\  
877: |\varkappa_t(t,x)|&\le&C_{t_0}\cdot t^{-\frac32}\label{A4}\\  
878: |\varkappa_x(t,x)|&\le&C_{t_0}\cdot t^{-\frac32}\label{A5}
879: \end{eqnarray}  
880: for all $t\ge t_0>0$ and all $x\in\mathbb T^1$.  
881: \end{corollary}  
882: \begin{proof}
883: We have already shown (\ref{A1}) and (\ref{A2}), explicitly. But on the
884: other hand estimates (\ref{A3}), (\ref{A4}) and (\ref{A5}) follow directly
885: from the above definition of~$\varkappa$ together with (\ref{A6}),
886: (\ref{Itom2}) and (\ref{Itom1}), respectively.
887: \end{proof}
888:  
889: \section{On the integration of constraints} 
890:  
891: As mentioned in the beginning part of this paper a special 
892: feature of Gowdy's model is the decoupling of the dynamical 
893: quantity~$a$. Furthermore, it is well-known that the 
894: momentum constraint 
895: \begin{eqnarray}   
896: \frac\partial{\partial x}a&=&\frac12\kern.5mm t \kern.5mm 
897: \frac\partial{\partial t}W\kern.5mm 
898: \frac\partial{\partial x}W \label{Impuls} 
899: \end{eqnarray}  
900: is conserved along $t$ developments; that means it is satisfied 
901: always if it is satisfied initially. 
902:  
903: With this remark in mind we will investigate now the    
904: Hamiltonian constraint 
905: \begin{eqnarray}  
906: \frac\partial{\partial t}a&=&-\frac1{4t}\kern1mm+  
907: \kern1mm\frac14\kern.5mmt 
908: \left[\left(\frac\partial{\partial t}W\right)^2 
909: \kern1mm+\kern1mm 
910: \left(\frac\partial{\partial x}W\right)^2\right] 
911: \label{Hamilton} 
912: \end{eqnarray}  
913: exclusively.  
914:   
915: \subsection*{Spatially homogeneous spacetimes} 
916: According to (\ref{Whom}) we find as the general solution 
917: of the Euler-Poisson-Darboux equation in spatially homogeneous 
918: spacetimes 
919: \begin{eqnarray*} 
920: W(t)&=&\beta\cdot\ln t\kern1mm+\kern1mm\gamma.  
921: \end{eqnarray*} 
922: The constraint (\ref{Hamilton}) reads 
923: \begin{eqnarray*}  
924: a_t(t)&=&\frac14(\beta^2-1)\cdot t^{-1},  
925: \end{eqnarray*}  
926: leading to 
927: \begin{eqnarray}  
928: a(t)&=&\frac14(\beta^2-1)\cdot\ln t\kern1mm+\kern1mm  
929: \zeta \label{ahom} 
930: \end{eqnarray}  
931: as solution where the arbitrary constant~$\zeta$ has to be 
932: specialized by an initial condition for $a$.  
933:   
934: \subsection*{Not spatially homogeneous spacetimes} 
935: The exceptional position of spatially homogeneous among 
936: the polarized Gowdy spacetimes is a consequence of the 
937: following consideration. 
938:  
939: \begin{theorem} 
940: Let $W\in\mathcal C^2(\mathbb R_+\times\mathbb T^1;\mathbb R)$ 
941: be a solution of the Euler-Poisson-Darboux equation. 
942: Then there is a positive constant~$C$, such that 
943: \begin{eqnarray*} 
944: |a_t(t,x)|&\le&C 
945: \end{eqnarray*} 
946: for all $t\in[t_0,\infty)$ and all $x\in\mathbb T^1$. 
947: \end{theorem} 
948:  
949: \begin{beweis} 
950: With corollary~\ref{Folgerung6} we have 
951: \begin{eqnarray*}  
952: W(t,x)&=&\gamma\kern1mm+\kern1mm\beta\cdot\ln t\kern1mm+\kern1mm  
953: t^{-\frac12}\nu(t,x)\kern1mm+\kern1mm\varkappa(t,x)  
954: \end{eqnarray*}  
955: with corresponding properties of $\nu$ and estimates for the
956: remainder~$\varkappa$; therefore
957: \begin{eqnarray}  
958: a_t&=&\frac14({\nu_t}^2+{\nu_x}^2)\kern1mm+\kern1mm\frac12  
959: \kern.5mm t^{-\frac12}\kern.5mm\beta\kern.5mm\nu_t     \nonumber\\ 
960: &&+\kern1mm\frac14 \kern.5mmt^{-1}\kern.5mm 
961: \left(\beta^2-1+2\nu_t\left(t^\frac32\varkappa_t-\frac12\nu\right) 
962: +2\nu_x t^\frac32\varkappa_x\right) \nonumber\\ 
963: &&+\kern1mm\mathcal O(t^{-\frac32}).                      \label{at}  
964: \end{eqnarray} 
965: The statement follows from corollaries~\ref{nu-Eigen}  
966: and~\ref{Folgerung6}. 
967: \end{beweis} 
968:  
969: \begin{remark} 
970: The above equation~(\ref{at}) already reflects the qualitatively 
971: different asymptotic behaviour in the metric function~$a$ depending on the 
972: spatial homogenity or not spatial homogenity of the underlying 
973: model. This is so because corollary~\ref{nu-Aequ} along with
974: corollary~\ref{om-Aequ}  
975: states that a polarized Gowdy spacetime is spatially 
976: homogeneous if and only if 
977: \begin{eqnarray*} 
978: \psi\kern1mm\equiv\kern1mm0&\leftrightarrow& 
979: \nu\kern1mm\equiv\kern1mm0\kern2mm\leftrightarrow\kern2mm 
980: \varkappa\kern1mm\equiv\kern1mm0. 
981: \end{eqnarray*} 
982: \end{remark} 
983:  
984: Since the asymptotics in spatially homogeneous models has been
985: completely investigated we will in what follows w.l.o.g.\ 
986: assume that   
987: $\nu_t\kern1mm=\kern1mm\nu_x\kern1mm=\kern1mm0$ is valid 
988: not everywhere. 
989:  
990: \begin{theorem} 
991: Let $a\in\mathcal C^2(\mathbb R_+\times\mathbb T^1;\mathbb R)$ 
992: be a solution of equation~(\ref{Hamilton}) and  
993: $\partial_x a\kern1mm=\kern1mm0$ not everywhere. Then there 
994: exist positive constants $\bar\nu$ and $C$ along with a uniquely 
995: determined function $\delta\in\mathcal C^2(\mathbb R_+\times 
996: \mathbb T^1;\mathbb R)$ such that 
997: \begin{eqnarray*} 
998: a(t,x)&=&\bar\nu\cdot t\kern1mm+\kern1mm\delta(t,x) 
999: \end{eqnarray*} 
1000: with 
1001: \begin{eqnarray*} 
1002: |\delta(t,x)|&\le&C\cdot (1+t^\frac12) 
1003: \end{eqnarray*} 
1004: for every $t\in[t_0,\infty)$, $t_0>0$ and every $x\in\mathbb T^1$. 
1005: \end{theorem} 
1006:  
1007: \begin{beweis} 
1008: Since $a$ is continously differentiable, there is in  
1009: $\mathbb R_+\times\mathbb T^1$ an open domain with 
1010: $a_x\kern1mm\neq\kern1mm0$ hence $\nu_x\kern1mm\neq\kern1mm0$. 
1011: Consequently 
1012: \begin{eqnarray}  
1013: \bar\nu&\mathop{=}\limits_{\rm Df}&\frac1{8\pi}  
1014: \intop_{t_0}^{t_0+2\pi}({\nu_t}^2+{\nu_x}^2)\kern1mm{\rm d}t  
1015: \label{nuq} 
1016: \end{eqnarray} 
1017: is a positive constant since it is also $x$-independent due to 
1018: $\nu_{tt}\kern1mm=\kern1mm\nu_{xx}$ and 
1019: \begin{eqnarray*}  
1020: \frac{\rm d}{{\rm d}x}\bar\nu&=&\frac1{4\pi}  
1021: \int_{t_0}^{t_0+2\pi}(\nu_t\nu_{tx}+\nu_x\nu_{xx})\kern1mm  
1022: {\rm d}t\\  
1023: &=&\frac1{4\pi}\int_{t_0}^{t_0+2\pi}\nu_x(-\nu_{tt}+\nu_{xx})  
1024: \kern1mm{\rm d}t\kern1mm+\kern1mm\left.\frac1{4\pi}\kern.5mm\nu_t  
1025: \kern.5mm\nu_x\kern.5mm\right|_{t_0}^{t_0+2\pi}\kern2mm=\kern2mm0. 
1026: \end{eqnarray*}  
1027: Let be further 
1028: \begin{eqnarray*}  
1029: \delta_t(t,x)&\mathop{=}\limits_{\rm Df}& 
1030: -\bar\nu\kern1mm+\kern1mm a_t(t,x)\\ 
1031: &=&-\bar\nu\kern1mm+\kern1mm\frac14({\nu_t}^2+{\nu_x}^2)\\ 
1032: &&+\kern1mm\frac12  
1033: \kern.5mm t^{-\frac12}\kern.5mm\beta\kern.5mm\nu_t \\ 
1034: &&+\kern1mm\frac14 \kern.5mmt^{-1}\kern.5mm 
1035: \left(\beta^2-1\kern.5mm+\kern.5mm 
1036: 2\nu_t\left(t^\frac32\varkappa_t-\frac12\nu\right) 
1037: \kern.5mm+\kern.5mm2\nu_x t^\frac32\varkappa_x\right) \\ 
1038: &&+\kern1mm\frac12\kern.5mmt^{-\frac32}\kern.5mm\beta\kern.5mm 
1039: \left(t^\frac32\varkappa_t\kern.5mm-\kern.5mm\frac12\nu\right) \\ 
1040: &&+\kern1mm\frac14\kern.5mmt^{-2}\left[\left(t^\frac32\varkappa_t- 
1041: \frac12\nu\right)^2\kern.5mm+\kern.5mm\left(t^\frac32\varkappa_x 
1042: \right)^2\right] 
1043: \end{eqnarray*} 
1044: which is uniquely determined for every fixed solution $W$  
1045: of the Euler-Poisson-Darboux equation. So we have 
1046: \begin{eqnarray*} 
1047: \delta(t,x)&=&\delta(t_0,x)\kern1mm+\kern1mm\frac14\int_{t_0}^t 
1048: \left({\nu_\tau}^2\kern.5mm+\kern.5mm{\nu_x}^2\right) 
1049: \kern1mm{\rm d}\tau\kern1mm-\kern1mm\bar\nu\cdot(t-t_0) 
1050: \kern1mm+\kern1mm\mathcal O(t^\frac12). 
1051: \end{eqnarray*} 
1052: Now we consider the second and third term of the 
1053: right hand side using~(\ref{nuq}) in detail: 
1054: \begin{eqnarray*}  
1055: \lefteqn{\mbox{}\kern-12mm\frac14\int_{t_0}^t\left({\nu_\tau}^2 
1056: \kern.5mm+\kern.5mm 
1057: {\nu_x}^2\right)\kern1mm{\rm d}\tau\kern1mm-\kern1mm 
1058: \bar\nu\cdot(t-t_0)}		\\[5mm] 
1059: \mbox{}\kern12mm&=& 
1060: \int_{t_0}^t\left[\frac14({\nu_\tau}^2+{\nu_x}^2)-\frac1{8\pi}  
1061: \intop_{t_0}^{t_0+2\pi}({\nu_t}^2+{\nu_x}^2)\kern1mm{\rm d}t  
1062: \right]\kern1mm{\rm d}\tau    \\[5mm] 
1063: \mbox{}\kern12mm&=& 
1064: \frac14\int_{t_0}^{t_0+2\pi\left[\frac{t-t_0}{2\pi}\right]}  
1065: ({\nu_t}^2+{\nu_x}^2)\kern1mm  
1066: {\rm d}t\kern1mm-\kern1mm\frac{\left[\frac{t-t_0}{2\pi}\right]}{4}  
1067: \intop_{t_0}^{t_0+2\pi}({\nu_t}^2+{\nu_x}^2)\kern1mm{\rm d}t\\[2mm] 
1068: \mbox{}\kern12mm&&+  
1069: \int_{t_0+2\pi\left[\frac{t-t_0}{2\pi}\right]}^t 
1070: \left[\frac14({\nu_\tau}^2+{\nu_x}^2)  
1071: \kern1mm-\kern1mm\frac1{8\pi}  
1072: \intop_{t_0}^{t_0+2\pi}({\nu_t}^2+{\nu_x}^2)\kern1mm{\rm d}t  
1073: \right]\kern1mm{\rm d}\tau  
1074: \end{eqnarray*}  
1075: Here as usual $[n]$ is the largest integer not 
1076: greater then $n$. Now, due to periodicity the first two terms cancel 
1077: each other. The integrand of the third is bounded, the domain 
1078: of integration less then $2\pi$ so it is  
1079: \begin{eqnarray*}  
1080: \left|\intop_{t_0}^t\left[\frac14({\nu_\tau}^2+{\nu_x}^2)  
1081: \kern1mm-\kern1mm\bar\nu  
1082: \right]\kern1mm{\rm d}\tau\right|&\le&2\pi C 
1083: \end{eqnarray*}  
1084: and finally 
1085: \begin{eqnarray*}  
1086: |\delta(t,x)|&\le&C\cdot(1+t^\frac12)  
1087: \end{eqnarray*}  
1088: for a positive constant $C$, all $x\in\mathbb T^1$  
1089: and every~$t\in[t_0,\infty)$, $t_0>0$. 
1090: \end{beweis} 
1091:  
1092: All results concerning the integration of constrains are 
1093: summarized by the following theorem.  
1094:   
1095: \begin{theorem}\label{Satz7}%  
1096: Let $W\in\mathcal C^2(\mathbb R_+\times\mathbb T^1;\mathbb R)$  
1097: be a solution of $W_{tt}+t^{-1}W_t-W_{xx}=0$.  
1098: Then any solution   
1099: $a\in\mathcal C^2(\mathbb R_+\times\mathbb T^1;\mathbb R)$ 
1100: of  
1101: \begin{eqnarray*}  
1102: a_t&=&-\frac14t^{-1}\kern1mm+\kern1mm\frac14t\kern.5mm  
1103: ({W_t}^2+{W_x}^2)  
1104: \end{eqnarray*}  
1105: may be cast for all $t\ge t_0>0$ and all $x\in\mathbb T^1$  
1106: into the form 
1107: \begin{eqnarray}  
1108: a(t,x)&=&\left\{  
1109: \begin{array}{ll}  
1110: \frac14(\beta^2-1)\cdot\ln t\kern1mm+\kern1mm\zeta,&  
1111: \text{if }\partial_x a(t,x)\equiv0\\[2mm]  
1112: \bar\nu\cdot t\kern1mm+\kern1mm\delta(t,x),&\text{otherwise}  
1113: \end{array}  
1114: \right. \label{aasym} 
1115: \end{eqnarray}  
1116: where $\bar\nu$ is a positive, $\zeta$ an arbitrary constant 
1117: and $\delta$ satisfies the inequalities 
1118: \begin{eqnarray*}  
1119: |\delta(t,x)|&\le&C\cdot (1+t^\frac12)\\[2mm]  
1120: |\delta_t(t,x)|&\le&C  
1121: \end{eqnarray*}  
1122: with another positive constante~$C$.  
1123: \end{theorem} 
1124:   
1125: \section{On geodesic completeness in polarized Gowdy spacetimes}  
1126:  
1127: In contrast to the analytical treatments in the first part 
1128: of this paper we will now investigate some open 
1129: geometrical questions in polarized Gowdy spacetimes. 
1130:  
1131: Isenberg and Moncrief~(1990) established important results 
1132: with respect to the singular behaviour in this class of  
1133: spacetimes. However, the question concerning possible 
1134: singularities in its future development remained open 
1135: (at least for the case with spatial $\mathbb T^3$-topology) and 
1136: shall be the subject of this section. As usual we will 
1137: identify the term ``singularity'' with the term 
1138: ``existence of an uncomplete non-spacelike geodesic'' 
1139: and use the basic concepts and relations of the theory 
1140: of Lorentzian manifolds and causality without repeating 
1141: all its definitions. However, for convenience we will summarize  
1142: some facts for short.  
1143:  
1144: Future will always mean the expanding direction 
1145: of the universe. 
1146:  
1147: In local coordinates 
1148: $(x^0,x^1,x^2,x^3)$ the geodesic equation reads: 
1149: \begin{eqnarray} 
1150: \frac{{\rm d}^2 x^i}{{\rm d}\tau^2}\kern1mm+\kern1mm 
1151: \sum_{j,k}\kern.5mm\Gamma^i_{jk}\kern.5mm 
1152: \frac{{\rm d} x^j}{{\rm d}\tau}\frac{{\rm d} x^k}{{\rm d}\tau} 
1153: &=&0,\kern10mmi,j,k=0,\dots,3\label{Geo} 
1154: \end{eqnarray} 
1155: Among the properties of affinly parametrized geodesics $\mathfrak x$
1156: the following are of special importance: 
1157:  
1158: 1.) The tangent vector is normalized 
1159: \begin{eqnarray} 
1160: \sum_{i,j}\kern.5mmg_{ij}\kern.5mm\frac{{\rm d}x^i}{{\rm d}\tau} 
1161: \kern.5mm\frac{{\rm d}x^j}{{\rm d}\tau}&=&-\varepsilon, 
1162: \label{Norm} 
1163: \end{eqnarray} 
1164: where $\varepsilon=0$, if the geodesic is null and 
1165: $\varepsilon=1$, if it is timelike. 
1166:  
1167: 2.) The scalar product of any Killing vectorfield $X$ and the 
1168: geodesic tangent is constant along the curve; in coordinates: 
1169: \begin{eqnarray} 
1170: \frac{{\rm d}}{{\rm d}\tau}\kern.5mm\sum_{i,j}\kern.5mmg_{ij} 
1171: \kern.5mm{X_\mathfrak x}^i\kern.5mm\frac{{\rm d}x^j}{{\rm d}\tau} 
1172: &=&0 
1173: \label{KVF} 
1174: \end{eqnarray} 
1175:  
1176: Gowdy spacetimes are known to be maximally extended, globally  
1177: hyperbolic regions which permit foliations by  
1178: $(t={\rm const})$-hypersurfaces. Consequently, any nonspacelike 
1179: curve intersects every such hypersurface exactly once. 
1180: This is important in view of the possibility to take 
1181: $t\to\infty$ limits along an arbitrary nonspacelike curve. 
1182: Furthermore, according to our choise of orientation it is  
1183: \begin{eqnarray} 
1184: \frac{{\rm d}t}{{\rm d}\tau}&>&0 \label{Orient} 
1185: \end{eqnarray} 
1186: if $\tau$ is as before the parameter of the future directed curve. 
1187:  
1188: Now we can start with the investigations. Since the asymptotic 
1189: expansion of polarized Gowdy spacetimes is quite different depending 
1190: on the spatial homogenity/inhomogenity of the model it is 
1191: necessary to split the proofs accordingly. 
1192:  
1193: \subsection*{Geodesics in spatially homogeneous, polarized 
1194: spacetimes}  
1195: Spatially homogeneous, polarized spacetimes correspond to 
1196: exactly solvable Einstein equations. We had found~(\ref{Whom})  
1197: and~(\ref{ahom}), 
1198: \begin{eqnarray}  
1199: W(t)&=&\gamma\kern1mm+\kern1mm\beta\cdot\ln t\label{Whom1}\\  
1200: a(t)&=&\zeta\kern1mm+\kern1mm\frac14(\beta^2-1)\cdot\ln t,  
1201: \label{ahom1} 
1202: \end{eqnarray}  
1203: with constants $\beta$, $\gamma$, $\zeta$. We need the 
1204: following estimate to prove geodesic completeness in such 
1205: models.  
1206:   
1207: \begin{lemma}  
1208: Let $\left(\mathbb R_+\times\mathbb T^3,g={\rm diag}(g_{00},g_{11},g_{22},g_{33})\right)$  
1209: be a spatially homogeneous solution of Einstein's field equation  
1210: for the polarized Gowdy model. Then there exist a constant~$C_\beta$, 
1211: such that the Christoffel symbols satisfy for every 
1212: $t\in[t_0,\infty)$:\label{Voll1} 
1213: \begin{eqnarray*}   
1214: \Gamma^0_{11}(t)&>&-\frac12 \kern1mm g^{00}\kern1mm 
1215: C_\beta\kern1mm t^{-1}\cdot g_{11} \\ [2mm] 
1216: \Gamma^0_{22}(t)&>&-\frac12 \kern1mm g^{00}\kern1mm  
1217: C_\beta\kern1mm t^{-1}\cdot g_{22} \\ [2mm] 
1218: \Gamma^0_{33}(t)&>& -\frac12\kern1mm g^{00}\kern1mm 
1219: C_\beta\kern1mm t^{-1}\cdot g_{33} 
1220: \end{eqnarray*}  
1221: \end{lemma}  
1222:   
1223: \begin{beweis}  
1224: The following section~\ref{Symbole} contains the relevant 
1225: Christoffel symbols. 
1226: Using~(\ref{Whom1}) and~(\ref{ahom1}) we find: 
1227: \begin{eqnarray*}  
1228: \Gamma^0_{11}(t)&=&-\frac12g^{00}\partial_t g_{11} 
1229: \kern2mm=\kern2mm-\frac12g^{00}\cdot\frac12(\beta^2-1)\kern1mm  
1230: t^{-1}\cdot g_{11}\\[2mm]   
1231: \Gamma^0_{22}(t)&=&-\frac12g^{00}\partial_t g_{22} 
1232: \kern2mm=\kern2mm-\frac12g^{00}\cdot\phantom{\frac12^2} 
1233: (\beta+1)\kern1mm 
1234: t^{-1}\cdot g_{22}\\ [2mm]  
1235: \Gamma^0_{33}(t)&=&-\frac12g^{00}\partial_t g_{33} 
1236: \kern2mm=\kern2mm-\frac12g^{00}\cdot\phantom{\frac12} 
1237: (-\beta+1)\kern1mm 
1238: t^{-1}\cdot g_{33} 
1239: \end{eqnarray*}   
1240: The constant $C_\beta$, 
1241: \begin{eqnarray} 
1242: C_\beta&\mathop{=}\limits_{\rm Df}&-\frac12(\beta^2+3),\label{Cbeta} 
1243: \end{eqnarray} 
1244: proves the statement. Observe that~$C_\beta<0$. 
1245: \end{beweis} 
1246:   
1247: \begin{theorem}  
1248: Every inextendable, future directed, nonspacelike geodesic in an 
1249: arbitrary but spatially homogeneous, polarized Gowdy manifold 
1250: is in its future direction complete.  
1251: \end{theorem}  
1252:   
1253: \begin{beweis}  
1254: It is sufficient to regard the $t$-component of an arbitrary 
1255: nonspacelike geodesic with affin parameter~$\tau$. Due to the 
1256: assumed homogenity not all of the symbols in section~\ref{Symbole} 
1257: are different from zero. More precisely, the $0$-component  
1258: of~(\ref{Geo}) reads 
1259: \begin{eqnarray*}   
1260: \frac{{\rm d}^2t}{{\rm d}\tau^2}\kern1mm+\kern1mm   
1261: \Gamma^0_{00}\left(\frac{{\rm d}t}{{\rm d}\tau}\right)^2\kern1mm+\kern1mm   
1262: \Gamma^0_{11}\left(\frac{{\rm d}x}{{\rm d}\tau}\right)^2\kern1mm+\kern1mm   
1263: \Gamma^0_{22}\left(\frac{{\rm d}y}{{\rm d}\tau}\right)^2\kern1mm+\kern1mm   
1264: \Gamma^0_{33}\left(\frac{{\rm d}z}{{\rm d}\tau}\right)^2&=&0. 
1265: \end{eqnarray*}   
1266: The causal character of the curve and the choise of time orientation 
1267: imply~(\ref{Orient}) and further 
1268: \begin{eqnarray}  
1269: \lefteqn{%\mbox{}\kern-4mm 
1270: \frac{\rm d}{{\rm d}t}\ln\left(\frac{{\rm d}t}{{\rm d}\tau}\right)^{-1} 
1271: \kern2mm=}\nonumber\\[2mm] 
1272: &&\kern-12mm\left(\frac{{\rm d}t}{{\rm d}\tau}\right)^{-2}\left[  
1273: \Gamma^0_{00}\left(\frac{{\rm d}t}{{\rm d}\tau}\right)^2\kern1mm+\kern1mm  
1274: \Gamma^0_{11}\left(\frac{{\rm d}x}{{\rm d}\tau}\right)^2\kern1mm+\kern1mm  
1275: \Gamma^0_{22}\left(\frac{{\rm d}y}{{\rm d}\tau}\right)^2\kern1mm+\kern1mm  
1276: \Gamma^0_{33}\left(\frac{{\rm d}z}{{\rm d}\tau}\right)^2  
1277: \right].\label{kern} 
1278: \end{eqnarray}  
1279: Now, using lemma~\ref{Voll1} we can estimate the symbols and get   
1280: \begin{eqnarray*}   
1281: \lefteqn{%  
1282: \Gamma^0_{11}\left(\frac{{\rm d}x}{{\rm d}\tau}\right)^2\kern1mm+\kern1mm   
1283: \Gamma^0_{22}\left(\frac{{\rm d}y}{{\rm d}\tau}\right)^2\kern1mm+\kern1mm   
1284: \Gamma^0_{33}\left(\frac{{\rm d}z}{{\rm d}\tau}\right)^2\kern2mm>\kern2mm}\\ 
1285: &&\kern14mm-\frac12 g^{00}  
1286: C_\beta \kern1mmt^{-1}\left[  
1287: g_{11}\left(\frac{{\rm d}x}{{\rm d}\tau}\right)^2\kern1mm+\kern1mm  
1288: g_{22}\left(\frac{{\rm d}y}{{\rm d}\tau}\right)^2\kern1mm+\kern1mm   
1289: g_{33}\left(\frac{{\rm d}z}{{\rm d}\tau}\right)^2  
1290: \right]. 
1291: \end{eqnarray*}  
1292: On the other hand we have~(\ref{Norm}) 
1293: \begin{eqnarray*}  
1294: g_{11}\left(\frac{{\rm d}x}{{\rm d}\tau}\right)^2\kern1mm+\kern1mm   
1295: g_{22}\left(\frac{{\rm d}y}{{\rm d}\tau}\right)^2\kern1mm+\kern1mm   
1296: g_{33}\left(\frac{{\rm d}z}{{\rm d}\tau}\right)^2&=&  
1297: -\varepsilon-g_{00}\left(\frac{{\rm d}t}{{\rm d}\tau}\right)^2 \\ 
1298: &<&-g_{00}\left(\frac{{\rm d}t}{{\rm d}\tau}\right)^2  
1299: \end{eqnarray*}  
1300: where $\varepsilon\in\{0,1\}$ so we get finally 
1301: ($C_\beta<0$)  
1302: \begin{eqnarray*}  
1303: \Gamma^0_{11}\left(\frac{{\rm d}x}{{\rm d}\tau}\right)^2\kern1mm+\kern1mm   
1304: \Gamma^0_{22}\left(\frac{{\rm d}y}{{\rm d}\tau}\right)^2\kern1mm+\kern1mm   
1305: \Gamma^0_{33}\left(\frac{{\rm d}z}{{\rm d}\tau}\right)^2  
1306: &>&\frac12 C_\beta \kern1mmt^{-1}  
1307: \left(\frac{{\rm d}t}{{\rm d}\tau}\right)^2.  
1308: \end{eqnarray*}   
1309: Using~(\ref{kern}) as well as~$\Gamma^0_{00}= 
1310: \frac14(\beta-1)\cdot t^{-1}$ and~(\ref{Cbeta}) 
1311: \begin{eqnarray*}   
1312: \frac{\rm d}{{\rm d}t}\ln\left(\frac{{\rm d}t}{{\rm d}\tau}\right)^{-1}  
1313: &>&\frac12  
1314: \left[   
1315: \frac12(\beta^2-1)+C_\beta  
1316: \right]\kern.5mm t^{-1}\kern2mm=\kern2mm-t^{-1}  
1317: \end{eqnarray*}   
1318: we get an suitable estimate. Integrating twice the result is 
1319: \begin{eqnarray*}  
1320: \tau(t)&>&\tau(t_0)\kern1mm+\kern1mmC\cdot\ln\frac{t}{t_0}  
1321: \end{eqnarray*}  
1322: for some positive $C$. Obviously, by taking $t\to\infty$ we have 
1323: proved that $\tau\to\infty$ holds along any nonspacelike 
1324: geodesic.  
1325: \end{beweis}  
1326:    
1327: \subsection*{Geodesics in not spatially homogeneous,  
1328: polarized spacetimes}  
1329:   
1330: In the case under consideration the components of Gowdy metric 
1331: are not everywhere $x$-independent and consequently, the 
1332: solutions of Einstein equations are not known explicitly. 
1333: Due to limitation on asymptotic behaviour as well as the 
1334: more complex form of geodesic equation proofs are more technical here.  
1335:   
1336: We begin by poving some estimates: 
1337:   
1338: \begin{lemma} 
1339: Let $t_0>0$ and $W\colon\mathbb R_+\times\mathbb T^1\to\mathbb R$  
1340: some smooth solution of $W_{tt}+t^{-1}W_t-W_{xx}=0$,  
1341: $W_x\equiv\kern-3.0mm/ \kern1.5mm0$. Then there is a positive  
1342: constant $C$, such that  for all $t>t_0$  
1343: \label{Voll2} 
1344: \begin{eqnarray*}  
1345: \left|-\frac1{2t^2}+\frac12({W_t}^2-{W_x}^2)\right|&\le&C\cdot t^{-1}  
1346: \end{eqnarray*}  
1347: holds. 
1348: \end{lemma}  
1349:   
1350: \begin{beweis} 
1351: For $t$ sufficiently large we know according to
1352: corollary~\ref{Folgerung6}  
1353: $W$ is of the form 
1354: \begin{eqnarray} 
1355: W(t,x)&=&\gamma\kern1mm+\kern1mm\beta\cdot\ln t\kern1mm+\kern1mm 
1356: t^{-\frac12}\left(\nu(t,x)\kern1mm+\kern1mm\omega(t,x)\right). 
1357: \label{W3} 
1358: \end{eqnarray}  
1359: Theorem~\ref{Satz4} proved boundedness of $\nu+\omega$ 
1360: while its partial derivatives are bounded by theorem~\ref{Satz5}, 
1361: hence the lemma.  
1362: \end{beweis}  
1363:   
1364: \begin{lemma}  
1365: For any polarized Gowdy spacetime model there is a constant~$C_1$  
1366: and a positive constant~$C$, such that  
1367: \label{Voll3} 
1368: \begin{eqnarray*}  
1369: |g_{ab}|\kern1mm+\kern1mm|\partial_t g_{ab}| 
1370: \kern1mm+\kern1mm|g^{ab}|\kern1mm+\kern1mm  
1371: |\partial_t g^{ab}|&\le&C\cdot t^{C_1}  
1372: \end{eqnarray*}  
1373: where $a,b\in\{2,3\}$ and $t_0<t<\infty$.  
1374: \end{lemma}  
1375:   
1376: \begin{beweis} 
1377: (\ref{W3}) shows logarithmic growth of $W$ if $\beta\neq0$, 
1378: consequently the metric coefficients $g_{ab}$, its inverses 
1379: and derivatives can increase at most with power~$\beta+1$.  
1380: So a constant~$C_1\mathop{=}\limits_{\rm Df}|\beta|+1$ 
1381: will be sufficient for our needs. 
1382: \end{beweis} 
1383:   
1384: \begin{theorem}  
1385: Every inextendable, future directed, nonspacelike geodesic in an 
1386: arbitrary but not spatially homogeneous, polarized Gowdy manifold 
1387: is in its future direction complete.  
1388: \end{theorem}  
1389:   
1390:   
1391: \begin{beweis}  
1392: Let $\mathfrak x=(t,x,y,z)$ be the components of some 
1393: arbitrary nonspacelike geodesic in local coordinates, 
1394: having affin parameter~$\tau$ and taking values in an 
1395: likewise arbitrary, polarized (not spatially homogeneous)  
1396: Gowdy manifold. As before it will sufficient to restrict 
1397: considerations on the $t$-component. Here it holds:  
1398: \begin{eqnarray}  
1399: 0&=&\frac{{\rm d}^2t}{{\rm d}\tau^2}\kern1mm+\kern1mm\Gamma^0_{00}  
1400: \left(\frac{{\rm d}t}{{\rm d}\tau}\right)^2\kern1mm+\kern1mm  
1401: 2\kern.5mm\Gamma^0_{01}\kern.5mm\frac{{\rm d}t}{{\rm d}\tau}  
1402: \frac{{\rm d}x}{{\rm d}\tau}\kern1mm+\kern1mm  
1403: \Gamma^0_{11}\left(\frac{{\rm d}x}{{\rm d}\tau}\right)^2\nonumber\\  
1404: &&+\kern1mm  
1405: \Gamma^0_{22}\left(\frac{{\rm d}y}{{\rm d}\tau}\right)^2\kern1mm+\kern1mm  
1406: \Gamma^0_{33}\left(\frac{{\rm d}z}{{\rm d}\tau}\right)^2 
1407: \label{kern2} 
1408: \end{eqnarray}  
1409: Again, Christoffel symbols are taken from section~\ref{Symbole}. 
1410: Using constraints~(\ref{Hamilton}) and~(\ref{Impuls})  
1411: we get: 
1412: \begin{eqnarray*}  
1413: \lefteqn{%  
1414: \mbox{}\kern-25mm 
1415: \Gamma^0_{00}\left(\frac{{\rm d}t}{{\rm d}\tau}\right)^2\kern1mm+\kern1mm  
1416: 2\kern.5mm\Gamma^0_{01}\kern.5mm\frac{{\rm d}t}{{\rm d}\tau}  
1417: \frac{{\rm d}x}{{\rm d}\tau}\kern1mm+\kern1mm  
1418: \Gamma^0_{11}\left(\frac{{\rm d}x}{{\rm d}\tau}\right)^2}\\[2mm]  
1419: \mbox{}\kern25mm&=&    
1420: \frac12 t\left(W_t\frac{{\rm d}t}{{\rm d}\tau}+   
1421: W_x\frac{{\rm d}x}{{\rm d}\tau}\right)^2   
1422: \kern1mm-\kern1mm\frac1{4t}  
1423: \left[\left(\frac{{\rm d}t}{{\rm d}\tau}\right)^2+   
1424: \left(\frac{{\rm d}x}{{\rm d}\tau}\right)^2\right]\\[2mm]  
1425: && -\kern1mm   
1426: \frac t4({W_t}^2-{W_x}^2)   
1427: \left[\left(\frac{{\rm d}t}{{\rm d}\tau}\right)^2-  
1428: \left(\frac{{\rm d}x}{{\rm d}\tau}   
1429: \right)^2\right]  
1430: \end{eqnarray*} 
1431: The first term on the right hand side is nonnegative and 
1432: will be neglected. Using condition~(\ref{Norm}) 
1433: in the form  
1434: \begin{eqnarray*}  
1435: \left(\frac{{\rm d}x}{{\rm d}\tau}\right)^2&=&  
1436: \left(\frac{{\rm d}t}{{\rm d}\tau}\right)^2  
1437: \kern1mm+\kern1mm g^{00} 
1438: \left[\varepsilon\kern.5mm+\kern.5mm  
1439: g_{22}\left(\frac{{\rm d}y}{{\rm d}\tau}\right)^2  
1440: \kern.5mm+\kern.5mm 
1441: g_{33}\left(\frac{{\rm d}z}{{\rm d}\tau}\right)^2\right]  
1442: \end{eqnarray*}   
1443: where as before $\varepsilon\in\{0,1\}$ depending on causal 
1444: character of geodesic, we get  
1445: \begin{eqnarray*}  
1446: \Gamma^0_{00} \left(\frac{{\rm d}t}{{\rm d}\tau}\right)^2\kern1mm+\kern1mm  
1447: 2\kern.5mm\Gamma^0_{01}\kern.5mm\frac{{\rm d}t}{{\rm d}\tau}  
1448: \frac{{\rm d}x}{{\rm d}\tau}\kern1mm+\kern1mm  
1449: \Gamma^0_{11}\left(\frac{{\rm d}x}{{\rm d}\tau}\right)^2 
1450: &\ge&-\frac1{2t}\kern.5mm  
1451: \left(\frac{{\rm d}t}{{\rm d}\tau}\right)^2		\\ 
1452: &&\mbox{}\kern-65mm+\left[-\frac1{4t}\kern.5mm+\kern.5mm  
1453: \frac t4({W_t}^2-{W_x}^2)\right]  
1454: \kern.5mm g^{00} \kern.5mm  
1455: \left[\varepsilon\kern.5mm+\kern.5mm  
1456: g_{22}\left(\frac{{\rm d}y}{{\rm d}\tau}\right)^2  
1457: \kern.5mm+\kern.5mm  
1458: g_{33}\left(\frac{{\rm d}z}{{\rm d}\tau}   
1459: \right)^2\right].  
1460: \end{eqnarray*}  
1461: This estimate yield for the factor  
1462: ($2t^{-1}$) of geodesic equation~(\ref{kern2}) 
1463: \begin{eqnarray}  
1464: \lefteqn{% 
1465: \frac{\rm d}{{\rm d}t}\left[t^{-1}\left(  
1466: \frac{{\rm d}t}{{\rm d}\tau}\right)^2\right]  
1467: \kern2mm\le}\nonumber\\[5mm] 
1468: &&\mbox{}\kern10mm 
1469: -g^{00}\left[-\frac1{2t^2}+\frac 12({W_t}^2-{W_x}^2)\right]  
1470: \left[\varepsilon\kern.5mm+\kern.5mm  
1471: g_{22}\left(\frac{{\rm d}y}{{\rm d}\tau}\right)^2\kern.5mm+\kern.5mm 
1472: g_{33}\left(\frac{{\rm d}z}{{\rm d}\tau}\right)^2  
1473: \right]\nonumber\\[2mm]  
1474: &&\mbox{}\kern10mm 
1475: +g^{00}\left[\partial_t  
1476: g_{22}\cdot t^{-1}\kern.5mm 
1477: \left(\frac{{\rm d}y}{{\rm d}\tau}\right)^2\kern.5mm+\kern.5mm  
1478: \partial_t  
1479: g_{33} 
1480: \cdot t^{-1}\kern.5mm\left(\frac{{\rm d}z}{{\rm d}\tau}\right)^2   
1481: \right],\label{Voll4} 
1482: \end{eqnarray}  
1483: where we have used~(\ref{Orient}). 
1484:  
1485: Due to~(\ref{KVF}) the scalar products~$g(\dot\mathfrak x,\cdot)$  
1486: in 
1487: \begin{eqnarray*} 
1488: \left(\frac{{\rm d}y}{{\rm d}\tau}\right)^2  
1489: &=&  
1490: \left[g^{22}\cdot g\left(  
1491: \dot\mathfrak x,\frac\partial{\partial y}\right)\right]^2\\[2mm]  
1492: \left(\frac{{\rm d}z}{{\rm d}\tau}\right)^2   
1493: &=&  
1494: \left[g^{33}\cdot g\left(  
1495: \dot\mathfrak x,\frac\partial{\partial z}\right)\right]^2  
1496: \end{eqnarray*}  
1497: are constant along the geodesic. 
1498: Thanks to lemmas~\ref{Voll2} and~\ref{Voll3} we can estimate 
1499: all brackets on the right hand side of the inequality~(\ref{Voll4})  
1500: by an $\mathcal O(t^C)$-term, where the constant~$C$ only depends 
1501: on the model parameter $\beta$ under consideration.  
1502: Furthermore, due to~(\ref{aasym}) it holds 
1503: $g^{00}\sim\mathcal O ({\rm e}^{-2\bar\nu t})$, $\bar\nu>0$, 
1504: $g^{00}$ will dominate the development on the right hand side.
1505: Consequently there are positive constants
1506: $C_1$, $C_2$ with
1507: \begin{eqnarray*}   
1508: \frac{\rm d}{{\rm d}t}\left[t^{-1}\left(   
1509: \frac{{\rm d}t}{{\rm d}\tau}\right)^2\right]   
1510: &\le&  
1511: C_1\kern.5mm{\rm e}^{-C_2t}  
1512: \end{eqnarray*}  
1513: and furthermore some positive constant 
1514: ${C_3}^{-2}$, such that for sufficiently large $t$ it holds  
1515: \begin{eqnarray*}  
1516: t^{-1}\left(  
1517: \frac{{\rm d}t}{{\rm d}\tau}  
1518: \right)^2&\le&{C_3}^{-2}  
1519: \end{eqnarray*}  
1520: and a final integration gives
1521: \begin{eqnarray*}  
1522: \tau(t)&\ge&\tau(t_1)\kern1mm+\kern1mm2{C_3}\left(\sqrt{t}-  
1523: \sqrt{t_1}\right),  
1524: \end{eqnarray*}  
1525: which proves $\tau\to\infty$ for $t\to\infty$.  
1526: \end{beweis}  
1527:   
1528: We summarize as follows:  
1529:   
1530: \begin{corollary}  
1531: Every inextendable, future directed, nonspacelike geodesic in an 
1532: arbitrary polarized Gowdy manifold 
1533: is in its future direction complete. 
1534: \end{corollary} 
1535:  
1536:  
1537: \section{Nonvanishing Christoffel symbols in polarized Gowdy spacetimes} 
1538: \label{Symbole} 
1539:  
1540: In polarized Gowdy models the nonvanishing components of the metric
1541: tensor are 
1542: \begin{eqnarray*}  
1543: g_{00}&=&-{\rm e}^{2a}\kern2cm g_{22}\kern2mm=\kern2mmt\kern.5mm{\rm e}^{W}\\  
1544: g_{11}&=&\phantom{-}{\rm e}^{2a}\kern2cm g_{33}\kern2mm=\kern2mmt\kern.5mm{\rm e}^{-W}  
1545: \end{eqnarray*}  
1546: where $a=a(t,x)$, $W=W(t,x)$, so up to symmetry the only nonvanishing
1547: symbols are  
1548: \begin{eqnarray*}  
1549: \begin{array}{lcrclclcrcl}  
1550: \Gamma^0_{00}&=&\frac12g^{00}\partial_t g_{00}&=&a_t&\phantom{mmm}&  
1551: \Gamma^0_{11}&=&-\frac12g^{00}\partial_t g_{11}&=&a_t\\[2mm] 
1552: \Gamma^0_{01}&=&\frac12g^{00}\partial_x g_{00}&=&a_x&\phantom{mmm}&  
1553: \Gamma^1_{01}&=&\frac12g^{11}\partial_t g_{11}&=&a_t\\[2mm] 
1554: \Gamma^1_{00}&=&-\frac12g^{11}\partial_x g_{00}&=&a_x&\phantom{mmm}&  
1555: \Gamma^1_{11}&=&\frac12g^{11}\partial_x g_{11}&=&a_x  
1556: \end{array}  
1557: \end{eqnarray*}  
1558: \begin{eqnarray*} 
1559: \begin{array}{lcrcl} 
1560: \Gamma^0_{22}&=&-\frac12g^{00}\partial_t g_{22}&=&-\frac12g^{00}  
1561: (t^{-1}+W_t)g_{22}\\[2mm] 
1562: \Gamma^2_{02}&=&\frac12g^{22}\partial_t g_{22}&=&\phantom{-}\frac12(t^{-1}+W_t)\\[2mm]  
1563: \Gamma^1_{22}&=&-\frac12g^{11}\partial_x g_{22}&=&-\frac12g^{11}  
1564: W_x g_{22}\\[2mm]  
1565: \Gamma^2_{12}&=&\frac12g^{22}\partial_x g_{22}&=&\phantom{-}\frac12W_x\\[2mm]  
1566: \Gamma^0_{33}&=&-\frac12g^{00}\partial_t g_{33}&=&-\frac12g^{00}  
1567: (t^{-1}-W_t)g_{33}\\[2mm]  
1568: \Gamma^3_{03}&=&\frac12g^{33}\partial_t g_{33}&=&\phantom{-}\frac12(t^{-1}-W_t)\\[2mm]  
1569: \Gamma^1_{33}&=&-\frac12g^{11}\partial_x g_{33}&=&\phantom{-}\frac12g^{11}  
1570: W_x g_{33}\\[2mm] 
1571: \Gamma^3_{13}&=&\frac12g^{33}\partial_x g_{33}&=&-\frac12W_x  
1572: \end{array} 
1573: \end{eqnarray*}  
1574:  
1575: %\begin{itemize} 
1576: %\item Interpretation: L"osungen der Einsteingleichungen 
1577: %korrespondieren zu linear polarisierten Graitationswellen mit 
1578: %$\mathbb T^3$-Topologie 
1579: %\item Isenberg und Moncrief (1990) untersuchten singul"ares 
1580: %Verhalten f"ur alle Topologien; $\mathbb S^3$ und  
1581: %$\mathbb S^2\times\mathbb S^1$ sind auch zukunftssingul"ar 
1582: %$\Rightarrow$ f"ur polarisierte Modelle blieb also nur offen, 
1583: %was in diesem Abschnitt abgehandelt wurde 
1584: %\item CIM machen Aussagen "uber nichtgenerische F"alle; 
1585: %Verbindung zu flachen Kasner-/Taubraumzeiten erw"ahnen; ggf.\ 
1586: %kurze Rechnung wie zu Misnerschen Raumzeiten; i.d.S.\ sind 
1587: %auch Bemerkungen im Schlu"sfolgerungsabschnitt von Moncrief (1981) 
1588: %sehr wertvoll $\Rightarrow$ ggf.\ m"ussen diese im obigen  
1589: %ausgeschlossen werden, was jedoch {\em sehr} unwahrscheinlich ist 
1590: %\end{itemize} 
1591:   
1592: %Das vorliegende Kapitel befa"ste sich mit der Struktur polarisierter  
1593: %Gowdymannigfaltigkeiten im Gro"sen. Der Schl"ussel dazu war  
1594: %die Absch"atzung von L"osungen  
1595: %der Euler-Poisson-Darboux-Gleichung. Die topologische  
1596: %Natur des Problems ausnutzend, h"atte man ebensogut mit einem  
1597: %Ansatz $W(t,x)=Z(t)R(x)$ beginnen und $R$ nach einem vollst"andigen  
1598: %System von Eigenfunktionen entwickeln k"onnen. Das Ergebnis  
1599: %ist eine formale L"osung  
1600: %\begin{eqnarray*}  
1601: %W(t,x)&=&\gamma+\beta\cdot\ln t+\sum_n^{1,\infty}  
1602: %b_n(J_0(nt)+N_0(nt))\cos nx +\\[-5mm]  
1603: %&&\phantom{\gamma+\beta\cdot\ln t+\sum_n^{1,\infty}}  
1604: %a_n(J_0(nt)+N_0(nt))\sin nx,   
1605: %\end{eqnarray*}  
1606: %in der $J_0$ und $N_0$ die Besselsche bzw.\ Neumannsche Funktion  
1607: %nullter Ordnung bezeichnen. Diese Reihe wurde bereits von   
1608: %in der Arbeit von Berger~(1974) angegeben.   
1609:   
1610: %F"ur unsere Zwecke bietet eine  
1611: %solch kurze Darstellung aber nur auf den ersten Blick  
1612: %Vorteile, da wir "uber die Anfangsbedingungen, die schlie"slich  
1613: %$a_n$ und $b_n$ bestimmen, damit aber auch "uber Konvergenzverhalten  
1614: %und Summe der Reihe zun"achst nichts wissen.  
1615:   
1616: %Besselsche Funktionen wurden zum ersten Mal von Schl"omilch~(1857)  
1617: %um ihrer selbst willen untersucht. Er betrachtete dabei auch  
1618: %Reihen des angegebenen Typs, die seither nach ihm benannt werden.  
1619: %Im Zusammenhang mit eindimensionalen Luftschwingungen gab  
1620: %Rayleigh~(1911) sp"ater eine physikalische Interpretation   
1621: %und von~Ignatowsky (1914, 1915) wurde etwa zur gleichen Zeit  
1622: %bei seinen Untersuchungen "uber Beugungsgitter auf dieselbe Reihe  
1623: %gef"uhrt, die er unabh"angig von den Vorgenannten   
1624: %untersuchte.~--~Jedenfalls ist bekannt   
1625: %(Twer\-sky (1961), Cooke (1928)),  
1626: %da"s sich nahezu  
1627: %beliebige Funktionen in Schl"omilchsche Reihen entwickeln lassen,  
1628: %was Aussagen "uber das asymptotische Verhalten von~$W$  
1629: %vermittels der angegebenen Reihe entsprechend erschwert.  
1630:   
1631: %\begin{itemize}  
1632: %\item Rayleigh- und Ignatkowskiverweise weglassen; letzten 
1633: %Satz umformulieren $\Rightarrow$ Reihe ist gliedweise L\"osung; 
1634: %Asymptotik der Besselschen Funktionen Richtlinie des Beweises 
1635: %\item Diskussion der Arbeit von Chru\'sciel, Isenberg und Moncrief  
1636: %\end{itemize} 
1637:   
1638: %Es sollte die Verbindung des vollen Modells / des polarisierten   
1639: %Modells zu den Kasnerl"osungen (insbesondere nat"urlich der  
1640: %r"aumlich homogenen L"osungen) ausgearbeitet werden.  
1641: %Die Rendall-Kichenassamy-Arbeit ist sicher kein ungeeigneter  
1642: %Ansatz dazu.  
1643:  
1644: %$\Rightarrow$ Asymptotisch homogen. In Abh\"angigkeit von gew\"ahlten 
1645: %Anfangsdaten werden alle Punkte mit Ausnahme eines Taubuniversums 
1646: %approximiert. Kurze Berechnungsskizze. Erw\"ahnung der Analogie von 
1647: %Isenberg/Moncrief f\"ur Kontraktion. 
1648:   
1649:  
1650: \section*{Acknowledgement} 
1651:  
1652: I am indebted to A.D.\ Rendall for suggesting me this problem and 
1653: for useful discussions. 
1654:  
1655: \section*{Literature}  
1656:   
1657: %{\bf Berger, B.K.} (1974) Quantum graviton creation in a model  
1658: %universe,  
1659: %{\em Ann. Phys.}, {\em 83}: 458-490.  
1660:   
1661: {\bf Chru\'sciel, P.T., J.~Isenberg and V.~Moncrief} (1990)   
1662: Strong cosmic censorship in polarised Gowdy spacetimes,  
1663: {\em Class. Quant. Grav.}, {\em 7}: 1671-1680.  
1664:   
1665: {\bf Gowdy R.H.} (1974)   
1666: Vacuum spacetimes with two-parameter spacelike isometry groups 
1667: and compact invariant hypersurfaces: topology and boundary 
1668: conditions,  
1669: {\em Ann. Phys.}, {\em 83}: 203-241.  
1670:  
1671: %{\bf Hawking, S.W.} (1968) The existence of cosmic time 
1672: %functions,  
1673: %{\em Proc. Roy. Soc. Lond.}, {\em A 308}: 433-435.  
1674:  
1675: %{\bf Hawking, S.W. and R. Penrose} (1970) The singularities of  
1676: %gravitaional collapse and cosmology,  
1677: %{\em Proc. Roy. Soc. Lond.}, {\em A 314}: 529-548.  
1678:   
1679: %{\bf Hopf, H. and W. Rinow} (1931) "Uber den Begriff der  
1680: %vollst"andigen differentialgeometrischen Fl"ache,  
1681: %{\em Comm. Math. Helv.}, {\em 3}: 209-225.  
1682:   
1683: %{\bf Isenberg, J. and V.~Moncrief} (1982)   
1684: %The existence of constant mean curvature foliations of Gowdy 
1685: %3-torus spacetimes,  
1686: %{\em Comm. Math. Phys.}, {\em 86}: 485-493.  
1687:  
1688: {\bf Isenberg, J. and V.~Moncrief} (1990)   
1689: Asymptotic behavior of the gravitational field and the nature  
1690: of singularities in Gowdy spacetimes,  
1691: {\em Ann. Phys.}, {\em 199}: 84-122.  
1692:    
1693: {\bf Moncrief, V.} (1981)   
1694: Global properties of Gowdy spacetimes with $\mathbb T^3\times\mathbb R$  
1695: topology, {\em Ann. Phys.}, {\em 132}: 87-107.  
1696:   
1697: {\bf Rendall, A.D.} (2000)   
1698: Fuchsian analysis of singularities in Gowdy spacetimes beyond analyticity,  
1699: {\em Class. Quant. Grav.}, {\em 17}: 3305-3316.  
1700:    
1701: {\bf Ringstr\"om, H.} (2002)   
1702: On Gowdy vacuum spacetimes,  
1703: {\em preprint}, {\tt gr-qc/0204044}.  
1704: 
1705: %{\bf Seifert, H.-J.} (1977)   
1706: %Smoothing and extending cosmic time functions,  
1707: %{\em Gen. Rel. Grav.}, {\em 8}: 815-831.  
1708:   
1709: \end{document} 
1710:   
1711: