1:
2:
3:
4:
5:
6: %\documentstyle[epsfig,prl,aps]{revtex}
7: %\documentstyle[prl,aps,twocolumn]{revtex}
8: %\documentstyle[preprint,eqsecnum,aps]{revtex}\def\btt#1{{\tt$\backslash$#1}}
9: \documentstyle[epsf,prl,aps,twocolumn]{revtex}
10: \begin{document}
11: \draft
12: \title{Gravitational redshifts in electromagnetic bursts occuring
13: near Schwarzschild horizon}
14: \author{ Janusz Karkowski $^1$ and}
15: \author{ Edward Malec $^{1,2}$ }
16: \address{ $^1$ Institute of Physics, Jagiellonian University, 30-064 Krak\'ow,
17: Reymonta 4, Poland }
18: \address{$^2$ Max Planck Institute for Gravitational Physics AEI, Golm, Germany }
19:
20:
21: \maketitle
22:
23: \begin{abstract}
24: It was suggested earlier that the gravitational redshift
25: formula can be invalid when the effect of the backscattering is strong.
26: It is demonstrated here numerically, for an exemplary electromagnetic pulse that
27: is: i) initially located very close to the horizon of a Schwarzschild black hole
28: and ii) strongly backscattered, that a mean frequency does not obey
29: the standard redshift formula. Redshifts appear to depend on
30: the frequency and there manifests a backscatter-induced blueshift in the
31: outgoing radiation.
32: \end{abstract}
33:
34:
35: \pacs{ 04.20.-q 04.70.-s 95.30.Sf 98.62.Js }
36: \date{ }
37:
38: \section{ Introduction}
39:
40:
41:
42: Standard derivations of the gravitational redshift base on
43: the approximation of geometric optics (see, for a discussion,
44: \cite{Sachs} - \cite{Strauman} ). This problem
45: has been recently reconsidered in the wave formulation
46: \cite{ME2002}. It was shown, for a class of
47: compact shocks that are separated in some sense \cite{sense}
48: from the horizon
49: of a Schwarzschild black hole, that the energy flux
50: scales accordingly to the redshift formula. Introducing
51: the concept of photons and assuming that their number is conserved,
52: one again arrives at the standard relation for the frequency.
53: As an alternative approach one can apply the Fourier
54: analysis, as outlined below in Sec. II.
55:
56: The crucial feature behind the above mentioned compactness and
57: separateness conditions is that when they hold true, then the backscatter
58: is negligible \cite{Karkowski}. It was remarked in \cite{ME2002} that
59: a compact pulse of radiation that is exposed to a significant
60: backscattering must not obey the familiar redshift relation. This
61: is suggested by the following reasoning. A spatially compact wave
62: packet consists of a superposition of monochromatic waves of different
63: frequencies. The effect of the backscattering is stronger for low
64: frequency waves than for high frequency ones, that is spectral
65: amplitudes of the former are stronger damped than those of the latter.
66: Therefore one expects that the backscattering induces a shift in a
67: mean frequency, so that it
68: does not satisfy the gravitational redshift formula. The aim of this
69: investigation is to extend results of \cite{ME2002} and find
70: a numerical example in favour of this conjecture. This examplary
71: wave pulse {\bf must break } -- and in fact {\bf does so},
72: in accordance with \cite{ME2002} -- the compactness condition of \cite{sense}.
73:
74: In the next section we write equations and review relevant
75: results of the preceding papers \cite{ME2002}, \cite{malec2000}.
76: Sec. III describes an electromagnetic wave that does not
77: comply to the standard redshift formula. Sec. IV presents a short summary.
78:
79:
80: \section{Gravitational redshift: classical derivation}
81:
82:
83:
84: The space-time geometry is defined by
85: the Schwarzschild line element,
86: %
87: \begin{equation} ds^2 =- (1-{2m\over R})dt^2 +
88: {1\over 1-{2m\over R}} dR^2 +
89: R^2 d\Omega^2~,
90: \label{1}
91: \end{equation}
92: %
93: where $t$ is a time coordinate, $R$ is an areal radius
94: and $d\Omega^2=d\theta^2 +\sin^2 \theta d\phi^2 $
95: (with $0\le \theta \le \pi $ and $0\le \phi <2\pi $)
96: is the line element on the unit sphere.
97: The Newtonian gravitational constant $G$
98: and the velocity of light $c$ are put equal to 1.
99: We define the tortoise coordinate $r^*=R+2m\ln ({R\over 2m}-1)$
100: and $\eta_R \equiv 1-{2m\over R}$.
101: We will consider only the dipole electromagnetic term and,
102: more specifically, choose the potential one-form $A=\sqrt{3/2}
103: \sin^2\theta \Psi (r^*,t)d\phi $.
104: The unknown function $\Psi $ satisfies the equation \cite{Wheeler}
105: %
106: \begin{equation}
107: (-\partial_t^2 + \partial_{r^*}^2)\Psi = \eta_R{2\over R^2}
108: \Psi .
109: \label{2}
110: \end{equation}
111: %
112: Let the electromagnetic strength field tensor be $F_{\mu \nu }$.
113: The stress-energy tensor of the electromagnetic field
114: reads $T_{\mu }^{\nu }=F_{\mu \gamma }
115: F^{\nu \gamma }-(1/4)g_{\mu }^{\nu }F_{\gamma \delta }
116: F^{\gamma \delta }$ and the time-like translational
117: Killing vector is denoted as $\zeta $. One can
118: define the energy flux $\hat P_R$,
119: %
120: \begin{eqnarray}
121: \hat P_R(R,t)\equiv {1\over \sqrt{\eta_r}}\int_{S(R)}dS(R)n_rT^{r\mu }\zeta_{\mu }.
122: \label{2.1}
123: \end{eqnarray}
124: %
125: Here $n$ is the unit normal to the sphere $R$ and $dS$ is the standard
126: area element on $S$.
127: $\hat P_R$ is equal to
128: %
129: \begin{eqnarray}
130: \hat P_R(R,t) =
131: -{4\pi \over \sqrt{\eta_R}} \partial_t\Psi \partial_{r^*}\Psi .
132: \label{2.5}
133: \end{eqnarray}
134: %
135: Let $\tilde \Gamma_{R_0}$ be a null geodesic directed outward
136: from the point $R_0$ of the initial hypersurface and let
137: $\tilde \Gamma_{ R_0, (R,t) }$ be a segment of $\tilde \Gamma_{R_0}$
138: that connects $R_0$ and $(R,t)$. Comparing the energy fluxes
139: through the spheres $S(R)$ (where $R>>2m$) and the initial $S(R_0)$,
140: one obtains \cite{ME2002}
141: %
142: \begin{eqnarray}
143: \hat P_R(R) \approx
144: \sqrt{\eta_{R_0}\over \eta_R} \hat P_R(R_0) ,
145: \label{2.2}
146: \end{eqnarray}
147: %
148: assuming that the initial support of a dipole wave packet is contained
149: in the annulus $(a,b)$ such that $(b-a)/(a\eta_a^5 )<<1$.
150: Here $a$ and $b$ are the {\bf areal radii}.
151:
152: There are two ways to derive the standard redshift
153: formula from (\ref{2.2}).
154:
155: i) {\it Eclectic approach}. The condition
156: %
157: \begin{equation}
158: (b-a)/(a\eta_a^5 )<<1
159: \label{cond}
160: \end{equation}
161: %
162: implies the validity of the geometric optics approximation.
163: Thence one can write $\hat P_R(R_0)= \hbar \hat \omega_{R_0} N_{R_0}$
164: and $\hat P_R(R)=\hbar \hat \omega_RN_{R}$, where $\hat \omega_{R_0}$
165: and $\hat \omega_R$ are the mean frequencies of the initial and
166: final pulses (as measured by static observers)
167: and $N_{R_0}$ and $N_R$ are the respective photon
168: fluxes. If the photon fluxes are conserved, then from (\ref{2.2})
169: one infers
170: %
171: \begin{equation}
172: \hat \omega_R=\sqrt{\eta_{R_0}\over \eta_R}\hat
173: \omega_{R_0}.
174: \label{2.4}
175: \end{equation}
176: %
177: ii) {\it Classical approach.} One can Fourier-analyze the quantity
178: representing the strength field tensor, $h=(-\partial_t+\partial_{r^*})\Psi $;
179: this is preserved along $\tilde \Gamma_{R_0}$, if condition (\ref{cond}) is satisfied.
180: Let the spectral strength field density be $g(\omega_{\tau } )=
181: \int dt e^{-i\omega_{ \tau }t} (-\partial_t+\partial_{r^*})\Psi $. Let the
182: support of $g(\omega )$ be $\tilde \Omega $. If one assumes the normalization
183: condition $\int_{\tilde \Omega } d\omega_{\tau } |g(\omega_{\tau })|=1$,
184: then the average frequency can be defined as $\hat \omega_{\tau } =
185: \int_{\tilde \Omega } d\omega_{\tau } \omega_{\tau } |g(\omega_{\tau })|$.
186: When the analysis is done with respect the asymptotic time $t$,
187: then asymptotic mean frequency $\hat \omega $ is found.
188: The Fourier analysis with respect
189: the proper time $d\tau = \sqrt{\eta_R}dt$ of a static observer located
190: at $R$ gives a corresponding mean frequency $\hat \omega /
191: \sqrt{\eta_R} $. Thus
192: the two frequencies satisfy (\ref{2.4}).
193:
194:
195:
196:
197:
198:
199: \section{Counterexample to the standard redshift formula}
200:
201:
202: It was pointed out in \cite{ME2002} that (\ref{2.4})
203: is valid for very compact initial pulses - those satisfying
204: the assumption (\ref{cond}). This condition
205: demands not only that a pulse is compact, but also that its relative width
206: is small in comparison to the relative distance from the black hole horizon.
207: It implies that the
208: approximation of the geometric optics is valid and that the
209: backscatter is absent. We conjecture, basing on a numerical evidence, that this
210: can be relaxed to $ (r^*(b)-r^*(a))/(2m)<<1$.
211:
212: In the case of a radiative pulse that is exposed to a significant
213: backscattering the relation (\ref{2.4}) would not hold.
214: Let us again review arguments in favour of this conclusion. A wave
215: pulse is superposed from monochromatic waves
216: of different frequencies. When resolved spectrally by an observer,
217: the pulse can be seen as a collection of peaks, each characterized by
218: some mean frequency. The effect of the backscattering is
219: more pronounced for low frequency waves than for high frequency ones
220: - the spectral amplitudes of the former are stronger damped than of
221: the latter. Therefore one would expect that mean frequencies
222: $\hat \omega_{R_0}, \hat \omega_{R}$ of the
223: initial and final pulses will not conform to (\ref{2.4}). The observed
224: mean frequency $\hat \omega_{R}$ of an outgoing pulse of radiation
225: can be blue-shifted in comparison to
226: the value $\sqrt{\eta_{R_0}\over \eta_{ R}}\hat \omega_{R_0}$.
227: In these circumstances an attempt to fit observed data
228: of a resolved multi-peak spectrum to the simple scaling law of
229: (\ref{2.4}) can lead to redshifts depending on the frequency.
230:
231: Our aim is to find numerical solutions corresponding to compact
232: initial data, satisfying the compactness assumption $b-a<<a$, such that:
233:
234: i) the spectral amplitude as a function of $\omega $
235: has several peaks with well resolved mean frequencies;
236:
237: ii) the evolution exhibits significant backscatter
238: (hence the energy flux is diminished);
239:
240: iii) some of mean frequencies at $R_0$ and $R$ do not comply
241: to (\ref{2.4}).
242:
243: The condition $b-a<<a$ (we remind the reader that $a$ and $b$ are areal radii)
244: is imposed in order to guarantee that initially
245: the wavelengths are much smaller than the radius of the wave front
246: and the curvature radius, i.e.,
247: that the assumption of the geometric optics is satisfied at the emission region.
248: On the other hand we {\bf do not assume} the validity
249: of (\ref{cond}); our initial data are such that $r^*(b)-r^*(a) = 4\pi m$,
250: which breaks down this inequality. Let us remark
251: that with such data one can have the geometric optics approximation
252: being valid both at the emission and detection regions, but being evidently
253: broken around $R\approx 3m$, when the wavelengths and the curvature
254: radius are of the same order.
255:
256: Below we present data describing one of many analysed examples.
257: The mass is normalized to unity and the Schwarzschild radius
258: $R_g$ is equal to 2. Let $I\equiv (r^*(a), r^*(b))$,
259: the support of initial radiation, be comprised between
260: $a=2+1.97\times 10^{-18}$ and $b= 2+10^{-15}$ or, in terms of the tortoise coordinate,
261: between $r^*(b)\approx -68.46$ and $r^*(a )\approx -68.46- 4\pi $.
262: The purely outgoing initial data have the form $\tilde \Psi = \partial_tf(r^*-t)
263: +f(r^*-t)/R$ \cite{malec2000}, where $f$ is a free datum in $I$ but vanishes
264: to the left from $a$.
265: With $I$ being so compact in terms of the areal radius it is reasonable
266: to ignore the $R$ - dependence in this expression and to prescribe
267: initial $\Psi $ as being dependent only on $r^*-t$.
268: Thus the initial data can be defined by prescribing $\tilde \Psi $
269: instead of $f$. We choose $\Psi (r^*)=\sin^2 ((r^*-r^*(a)))$ and
270: $\partial_t\Psi (r^*)=-\sin (2(r^*-r^*(a)))$ for
271: $r^*\in I\equiv (r^*(a), r^*(b))$ and $\Psi (r^*)=\partial_t\Psi
272: =0$ outside $I$.
273:
274: Notice that the size of the support of the initial pulse, $i=4\pi $, satisfies
275: the inequality $i>2$, which is the plausibility condition (according to
276: the foregoing conjecture) for a significant backscatter.
277:
278: The signal is measured during the time
279: interval $32 \pi $ at two observation
280: points, $r^*_{0}=-68.44$ and $r^*_1= 199.16$; in terms of the
281: areal radius we have $R_0=2+1.012\times 10^{-15}$ and $R=190.07$.
282: The grid was $120000\times 60000$.
283: Two snapshots of the evolving configuration taken at the
284: above pair of observation
285: points are presented in Fig. 1.
286: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
287: \begin{figure}[1]
288: \epsfxsize=6cm
289: \centerline{\epsffile{j2.eps}}
290: \caption{ Temporal dependence of $h\equiv (-\partial_t+\partial_{r^*})\Psi $
291: (with values of $h$ put on the ordinate)
292: as observed at $R_0=(2+10^{-15})m$ (solid line) and at $R=190.07m$ (broken line).
293: The time (with values put on the abscissa) is measured from the moment
294: of the arrival of the wave front at each of the
295: observation points. The part of the radiation seen at $R$ (broken line)
296: after $t=4\pi m$ consists solely of the backscattered radiation.
297: The signal is negligible after $t>30m$ - its amplitude is much smaller than that
298: of the oscillatory part.}
299: \end{figure}
300: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
301:
302: The main pulse passes through $R$ in less than
303: $4\pi $ (in units of $m$) while the scattered wave appears later and
304: quickly vanishes, becoming negligible after $t=30 m$, when the tail dominates.
305: The decay of the tail asymptotics should be like $t^{-5}$,
306: according to \cite{Price}. This asymptotic exponent was in fact obtained
307: in many trial runs of our numerical code with initial pulses
308: located outside $a=3m$ \cite{expla}.
309:
310:
311:
312: Figure 2
313: in turn shows the spectral decomposition of $h=(-\partial_t+
314: \partial_{r^*})\Psi $, obtained by employing the FOURCO package
315: (netlib). Fast Fourier Transform (IMSL library) was also used,
316: with the purpose of checking the numerical results coming from the
317: application of FOURCO. We depicture
318: {\bf normalized frequencies} in order to see clearly the frequency shift
319: that is caused by the backscatter. The spectrum that is seen at $R_0 $
320: (solid line) is rescaled by $\sqrt{\eta_{R_0}}$ and the
321: spectrum seen at $R$ (broken line) is rescaled by the factor
322: $\sqrt{\eta_R}$. In the case of negligible
323: backscattering both rescaled spectra would have to coincide.
324:
325: Figure 2 shows the spectral amplitudes $|g(\omega )|$ in function of the
326: frequency $\omega $. There is a number of peaks, seen at
327: both observation points. We calculated average frequency for
328: each of the resolved peaks, by choosing as the support $\tilde \Omega $
329: of each pulse the interval between the consecutive minima of
330: $|g(\omega )|$. In the case of the lowest frequency peak one can observe
331: a 25-percent blueshift, from 0.25 to 0.31, with variances 0.1
332: and 0.08, respectively; the broken and solid lines of Fig. 2
333: coincide for $\omega >0.5/m$. In the remaining peaks the effect is
334: seemingly absent. This can be interpreted as demonstrating that this
335: backscatter-induced blueshift is frequency dependent.
336: $1/(4m)$ is approximately the asymptotic frequency of the quasinormal modes.
337: It can be seen from Fig. 2 that the effect of the
338: backscattering strongly weakens
339: modes having frequencies $\omega_b$ (as observed by an observer
340: static at $b$), such that $\omega_a\sqrt{\eta_b}$ (which
341: would be the frequency detected by an asymptotic static observer)
342: is smaller than $1/(8m)$ - a half of the asymptotic frequency of quasinormal modes.
343: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
344: \begin{figure}[2]
345: \epsfxsize=6cm
346: \centerline{\epsffile{j1.eps}}
347: \caption{Frequency amplitude $|g(\omega )|$ (depictured on the ordinate)
348: as observed at $R_0=(2+10^{-15})m$ (solid line) and at $R=190.07m$ (broken line).
349: The {\bf normalized frequencies} (see the text above)
350: are put on the x-axis (scaled in units of $1/m$). }
351: \end{figure}
352: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
353:
354:
355: Modes with $ \omega_b\sqrt{\eta_b} >1/(2m)$ are unchanged, while those
356: in the interval $(1/(8m)<\omega_b\sqrt{\eta_b}< 1/(2m)$
357: undergo a damping of the spectral amplitude,
358: which is the stronger the smaller the frequency. The physical reason
359: for this is following. Our initial wave pulse has to penetrate through the
360: potential barrier having maximum at $r=3m$; the modes with frequencies
361: close to those of quasinormal modes are much stronger scattered and their
362: transmission coefficient is small.
363:
364: Let us point out that the backscatter is negligible -- due to the
365: weakness of the effective potential --
366: as long as a wave pulse remains in a region
367: characterized by the tortoise variable $r^*<<-1$.
368: Therefore the standard redshift formula applies in the
369: zone $r^*<<1$.
370: Define a shifted radiation pulse as follows: $\Psi (r^*+D)=\Psi(r^*)$
371: for $D<0$). Notice that this translated pulse will possess
372: a {\bf different} frequency profile. Let $b_D$ be defined by the equation $r^*(b)+D=
373: b_D+2\ln (b_D/(2m)-1)$ and let $(w_b)$ be a frequency measured by a
374: static observer located at $r^*(b)$.
375: A static observer located at $r^*(b)+D$ will measure a blueshifted frequency
376: $(w_b)\sqrt{\eta_b/\eta_{b_D}}$.
377: On the other hand {\bf a static observer located at infinity
378: would detect the same set of frequencies irrespective of $D$}.
379: Moreover, this observer would notice that the
380: deviation from the redshift formula (if that can be seen in the detected
381: frequency range) does not depend on the position of the initial pulse.
382:
383: If however a {\bf physical} source of a wave
384: is fixed and shifted
385: from $r^*(b) <<-1$ by some $D<0$ then
386: the two local sets of frequencies $(\omega_{bi}, \omega_{b_Di}$)
387: measured by two (shifted by a distance $D $) local static observers {\bf do coincide},
388: $\omega_{bi}=\omega_{b_Di}$. (We assume that tidal effects can be ignored.)
389: The static observer at infinity would then see:
390: i) two {\bf different sets of redshifted frequencies}, $(\omega_{bi,\infty })$
391: and $(\omega_{b_Di,\infty })$, with $\omega_{bi,\infty }=\sqrt{\eta_b}\omega_{bi}$
392: and $\omega_{b_Di,\infty }=\sqrt{\eta_{b_D}}\omega_{b_Di}$
393: for all frequencies which satisfy conditions
394: $\omega_{bi,\infty }>>1/(4m) $ and $\omega_{b_Di,\infty }>>1/(4m)$;
395: ii) those frequencies, for which
396: $\omega_{bi,\infty }\le 1/(4m)$ and/or $\omega_{b_Di,\infty }\le 1/(4m)$, would
397: disobey the standard redshift formula. It is clear that for any fixed wave
398: source with frequencies $\omega_i$ one can find a location $b$ close to the
399: black hole horizon such that some of
400: (the would-be asymptotically detected) frequencies $\eta_b\omega_i$ are of the order
401: of $1/(4m)$, so that (according to our numerical example) one could see a failure
402: of the standard redshift formula.
403:
404: If that numerically discovered phenomen is generic, then there would
405: exist a natural cutoff (of the order of $1/(4m)$) for the asymptotically
406: observed frequency of those wave pulses that are close to an event horizon.
407: A wave of a mean frequency $\omega_b$
408: satisfying the condition $\omega_b\sqrt{\eta_a}<< 1/(4m)$
409: would not be observed by a distant observer.
410: In the light of the above the heuristic
411: analysis that was reported in the
412: beginning of this section, should be probably supplemented by adding
413: the following: {\it if a wave peak possesses a contribution
414: with frequencies much smaller than $1/(4m)$ then (and only then) its
415: mean frequency will be influenced by the backscatter}.
416:
417:
418:
419:
420:
421:
422:
423:
424: \section{Final remarks}
425:
426: The backscattering can modify the relation between initial and asymptotic
427: frequencies. A characteristic frequency (circa $1/(4m)$ in the hitherto
428: used units) that appears useful in this
429: context is (in standard units) $f_c\approx 0.25\times 31484\times M_0/m$ Hz,
430: where $M_0$ is the solar mass. Let us remark that $f_c$ is the
431: frequency typical for the quasinormal modes of the electromagnetic
432: radiation propagating in the Schwarzschild spacetime.
433:
434: A numerical example of Sec. III reveals a robust difference between
435: the standard redshift prediction (\ref{2.4}) and the actual observation.
436: It is observed that for a radiation pulse having an asymptotic frequency
437: $f_{\infty }\approx f_c$ the initial frequency $f_R$ is
438: noticeably smaller than $f_c/\sqrt{\eta_R}$, a number following
439: from the redshift formula (\ref{2.4})). The peaks with frequencies $f_{\infty }>>f_c$
440: are seemingly not influenced by the backscattering. This probably
441: is not always true for all radiation pulses having mean frequencies
442: much bigger than $f_c$,
443: but we believe that in the latter case the deviation from the redshift
444: formula must be small. Rephrasing this fact in terms of initial
445: data, it is probably fair to say that $(r^*(b)-r^*(a)<<2m$ is
446: the sufficient condition for the validity of the redshift formula (\ref{2.4}).
447: That is a much less stringent condition than that used in
448: \cite{{ME2002}}.
449: Let us stress that in the case of astrophysical black holes of stellar
450: origin the deviation from the law (\ref{2.4}) can be seen
451: by an asymptotic observer in the radio part of the electromagnetic
452: spectrum, with wavelengths of the order of 100 km; we do not claim that
453: this phenomenon is of astrophysical interest.
454:
455: Let us point out that if the backscattering is negligible then:
456: i) all of the energy of an outgoing pulse
457: would get to infinity; ii) its energy flux would be diminished,
458: according to (\ref{2.5}). The only mechanism that can imply
459: the loss of the radiation energy is through the backscatter of the
460: radiation off the curvature of the spacetime. This is
461: clearly shown in the wave formulation \cite{malec2000}, but a consistent
462: quantum field theoretic treatment (in the curved Schwarzschild spacetime)
463: should give the same answer.
464:
465: In the present paper the consideration is focused only on the dipole
466: term, but a similar analysis with the same conclusions can
467: to be done in any multipole order.
468: An analogous phenomenon, leading to the demodulation of gravitational
469: signals, will also manifest in the spectra of gravitational waves;
470: this in principle may be detected.
471:
472: Acknowledgments. This work has been suported
473: in part by the KBN grant 2 PO3B 006 23.
474:
475:
476:
477: \begin{references}
478: \bibitem{Sachs} J. Kristian and R. K. Sachs, {\it Astrophysics Journal}
479: {\bf 143}, 379(1965).
480: \bibitem{Ehlers} J. Ehlers, "Survey of general relativity theory",
481: in: Relativity, Astrophysics and Cosmology, W. Israel (ed.), Reidel
482: Publ. Comp., Dordrecht,
483: p.1 1973.
484: \bibitem{Schneider} P. Schneider, J. Ehlers and E. E. Falco,
485: Gravitational Lenses,
486: Springer-Verlag, Berlin, Heidelberg, New York, 1992.
487: \bibitem{Strauman} N. Straumann, General Relativity and Relativistic
488: Astrophysics, Springer Verlag 1984.
489: \bibitem{ME2002} E. Malec, {\it Class. Quantum Grav. } {\bf 19}, 571(2002).
490: \bibitem{sense} It is meant here that $(b-a)/(a(1-2m/a)^5 )<<1$,
491: where the areal radii $a$ and $b$ are the inner and outer boundaries
492: of the initial wave pulse.
493: \bibitem{malec2000} E. Malec, {\it Phys. Rev.} {\bf D62}(2000), 084034.
494: \bibitem{Karkowski} J. Karkowski, E. Malec and Z. \'Swierczy\'nski,
495: {\it Class. Quantum Grav. } {\bf 19}, 953(2002).
496: \bibitem{Wheeler} J. A. Wheeler, {\it Phys. Rev.} {\bf 97} 511.
497: \bibitem{Price} R. Price, {\it Phys. Rev. } {\bf D5}, 2419(1972).
498: \bibitem{expla} The integration time, that would be needed in order to obtain the
499: asymptotic tail behaviour in the case of the initial data of Sec. III,
500: exceeds by far the integration time required in order to solve our task.
501:
502: \end{references}
503:
504:
505:
506:
507:
508: \end{document}
509:
510:
511:
512:
513:
514:
515: