1: % -*-latex-*-
2: \documentclass[prl,final,aps,twocolumn]{revtex4}
3:
4: %\documentstyle[aps,prl,epsfig]{revtex}
5: %\documentstyle[aps]{revtex}
6:
7: \usepackage{graphicx}
8:
9: \begin{document}
10:
11:
12: %\bibliographystyle{prsty}
13:
14:
15:
16:
17: \title{The intermediate problem for binary black hole
18: inspiral and the periodic standing wave approximation
19: }
20:
21:
22:
23:
24: \author{Benjamin Bromley}
25: %
26: \affiliation{Department of Physics,
27: University of Utah, Salt Lake City, Utah 84112}
28: %
29: \author{Robert Owen}
30: \affiliation{Theoretical Astrophysics, California
31: Institute of Technology, Pasadena, CA 91125}
32: %
33: \author{Richard H.~Price}
34: \affiliation{Department of Physics \& Astronomy and Center for
35: Gravitational Wave Astronomy, University of Texas at Brownsville,
36: Brownsville, TX 78520}
37:
38:
39: \begin{abstract}
40: {\bf Abstract}
41: In calculations of the inspiral of binary black holes an intermediate
42: approximation is needed that can bridge the post-Newtonian methods
43: of the early inspiral and the numerical relativity computations of
44: the final plunge. We describe here the periodic standing
45: wave approximation: A numerical solution is found to the problem
46: of a periodic rotating binary with helically symmetric standing
47: wave fields, and from this solution an approximation is
48: extracted for the physically relevant problem of inspiral with outgoing waves.
49: The approximation underlying this approach has been recently confirmed
50: with innovative numerical methods applied to nonlinear model problems.
51:
52:
53: \end{abstract}
54:
55:
56: \maketitle
57:
58:
59: The computational study of the inspiral of binary black holes is
60: important for the understanding of gravitational wave signals, and is
61: of inherent interest as a question in general relativity that can be
62: answered only with computation. It has therefore become the focus of
63: supercomputer codes that evolve Einstein's field equations forward in
64: time from initial conditions chosen to represent a starting
65: configuration of the inspiralling objects. The evolution codes,
66: however, typically become unstable on a timescale (set by the size of
67: the hole) short compared to a full orbit. Reliable calculations of the
68: final plunge are now feasible, the merger and ringdown of the final black
69: hole fate of the system are handled well with perturbation
70: theory\cite{laz1}, and the early inspiral is well approximated with
71: post-Newtonian computations\cite{PN}. What cannot be handled well is
72: the intermediate phase of the inspiral, that late epoch during which
73: nonlinear effects are too strong for the post-Newtonian approximation,
74: and too many orbits remain for numerical relativity to be stable.
75:
76: It has long been recognized that the basis of an approximation scheme
77: should be the slow rate of inspiral, the small ratio of the orbital
78: time to the radiation damping time\cite{det,det2}. Through an
79: adiabatic treatment of the slow inspiral, such an approximation could
80: give answers about the radiation and rate of inspiral in the
81: intermediate epoch. And when the rate of inspiral became too rapid,
82: the intermediate approximation could hand the problem off to numerical
83: evolution codes to do the final orbit and plunge, and could supply the
84: ideal initial data to those evolution codes. The need and the concept
85: for an intermediate approximation have been clear, but such an
86: approximation has not been easy to implement. Along with several
87: colleagues\cite{WKP,WBLandP,rightapprox,paperI} we have based an
88: approximation of slow inspiral on a numerical computation of no
89: inspiral. That is, we seek a numerical solution of Einstein's
90: equations for binary objects that are in circular periodic motion, and
91: whose ``helically symmetric'' fields rotate rigidly with the source
92: objects.
93:
94: The universality of gravitation suggests that the unchanging motion of
95: such a system is not compatible with outgoing radiation, and this
96: intuitive suggestion is confirmed by the mathematics of the theory.
97: We, therefore, seek a helically symmetric solution for the sources
98: coupled to standing waves. In a linear theory standing waves would be
99: a superposition of half-ingoing and half-outgoing solutions. In
100: linear theory, one could (though without motivation) solve the
101: standing wave problem. From the fact that solution is half the
102: superposition of the ingoing and outgoing solutions, and from the
103: relationship of the ingoing and outgoing solutions, one could extract
104: the outgoing solution. The crux of our periodic standing wave
105: method is that even for highly nonlinear binary inspiral fields there
106: is an ``effective linearity.'' The standing wave solution, to good
107: accuracy, is half the sum of the outgoing plus ingoing solutions
108: despite the nonlinearities. In general relativity, therefore, we
109: should be able to solve the standing wave numerical problem and extract an
110: approximation to the outgoing solution.
111:
112: It is important to understand why effective linearity can be correct
113: for inspiral. In the strong-field regions very close to the sources,
114: the solution is very insensitive to the distant radiative
115: boundary conditions (ingoing, outgoing, standing wave). In this
116: near-source region a superposition of half the ingoing and half the
117: outgoing solution gives a good approximation solution, because
118: it amounts to averaging two samples of the same thing. In the wave zone where
119: the outgoing and the ingoing solutions are very different, the fields
120: are weak enough that nonlinear effects are negligible, and once again
121: we can superpose. The separation of the strong-field region from the
122: boundary-influenced region should be a clear separation unless the
123: sources are rotating very close to $c$, in which case the wave zone
124: will start just outside the sources. It is, however, not expected
125: that ultrarelativistic source motion can occur during the slow
126: inspiral epoch of motion.
127:
128: We have recently been able to confirm effective linearity. (Technical
129: details are given in \cite{eigenspec}.) This confirmation has been
130: achieved with a model problem, since the validity of effective linearity
131: can only be carried out in a model problem. In general relativity,
132: there will be no ``true outgoing'' solution available for confirmation
133: until numerical evolution codes are fully developed. In addition, the
134: numerical features of the helically symmetric standing wave
135: calculation pose new challenges very different from those of evolution
136: codes, and are best resolved in the simplest context possible.
137:
138:
139:
140:
141:
142: Our model problem is a nonlinear scalar field coupled to point-like
143: sources in Minkowski space, and satisfying the field equation
144: \begin{equation}\label{fieldtheory}
145: \Psi_{;\alpha;\beta}g^{\alpha\beta}
146: +\lambda F=\nabla^2\Psi-\frac{1}{c^2
147: }\partial_t^2\Psi+ F={\rm Source}\,.
148: \end{equation}
149: Here the source is taken to be two points of unit scalar charge in
150: orbit around each other at angular frequency $\Omega$, and at radius
151: $a$. The velocity parameter for the system $\beta=a\Omega/c$ is taken
152: to be of order unity, representing the strong-field tight binary for
153: which post-Newtonian approximations are inadequate. Our methods,
154: however, require that $\beta$ not be too close to unity. More
155: explicitly, our method works best if the radiation is quadrupole
156: dominated, and our calculations focus on values of $\beta $ from 0.3
157: to 0.5. We expect that the approximation of slow inspiral will break
158: down before this assumption breaks down, and numerical relativity
159: evolutions will take over the job of tracking the last part of an
160: orbit and the subsequent plunge and merger.
161:
162: The term $F$ contains the nonlinearity in our model theory, and we
163: have found the following form, with parameters $\lambda$ and $\Psi_0 $,
164: to be very useful:
165: \begin{equation}\label{modelF}
166: F=\frac{\lambda}{a^2}
167: \frac{\Psi^5}{\Psi_0^4+\Psi^4}\,.
168: \end{equation}
169: A crucial feature of $F $ is that like the nonlinearities of general
170: relativity, it is very large near the sources, and becomes negligible
171: far from the sources. The $\lambda$ multiplier allows us to vary the
172: strength of the nonlinear term, and the $\Psi_0$ parameter allows us
173: to vary the profile of the nonlinearity in the strong field region.
174:
175:
176:
177: %
178: Our problem is defined by Eqs.~(\ref{fieldtheory}) and (\ref{modelF}),
179: and the source motion at angular frequency $\Omega$ in the equatorial
180: plane. As described in spherical coordinates, helical symmetry can be
181: imposed on the solution $\Psi(t,r,\theta,\phi)$ by restricting to
182: solutions of the form $\Psi(r,\theta,\varphi) $, where $\varphi $ is
183: the comoving azimuthal coordinate $\phi-\Omega t $. By restricting
184: the solution in this way, we have eliminated the possibility of
185: ``evolution.'' For such helically symmetric solutions a change in
186: time by $\Delta t$ is the same as a change the azimuthal angle
187: $\Delta\phi=-\Omega\Delta t$. With this suppression of evolution, we
188: have eliminated the sorts of instabilities that develop in evolution
189: codes. But we have introduced new difficulties.
190:
191: These new difficulties can be seen immediately in the form of the
192: helically restricted nonlinear scalar equation
193:
194: %\begin{equation}\label{ourprob}
195: %{\cal L}(\Psi)=\sigma_{\rm eff}[\Psi]\,,
196: %\end{equation}
197: %with
198: %${\cal L}$ taken to be
199: \begin{displaymath}
200: {\cal L}\Psi\equiv
201: \frac{1}{r^2}\frac{\partial}{\partial r}
202: \left(r^2\frac{\partial\Psi}{\partial r}\right)
203: +\frac{1}{r^2\sin\theta}\frac{\partial}{\partial\theta}
204: \left(\sin\theta\frac{\partial\Psi}{\partial\theta}\right)
205: \end{displaymath}
206: \begin{equation}\label{ourL}
207: +\left[\frac{1}{r^2\sin^2\theta}
208: -\frac{\Omega^2}{c^2}
209: \right]\,\frac{\partial^2\Psi}{\partial\varphi^2}
210: ={\rm Source}-F(\Psi)\equiv\sigma(\Psi)\,.
211: \end{equation}
212: The principal part of this quasilinear equation is ``mixed,'' elliptic
213: inside a cylinder at $r\sin\theta=c/\Omega $, and hyperbolic outside
214: that cylinder. The problem is to be solved with radiative conditions
215: (ingoing, outgoing, or standing wave as described below) on a
216: spherical surface at large distances. Well posed problems in physics
217: typically supply cauchy data on open surfaces to hyperbolic equations,
218: and Dirichlet or Neumann data on closed surfaces to elliptic
219: equations. Our model leads to a boundary value problem with
220: ``radiation'' conditions on a closed surface surrounding a mixed
221: problem. Though unusual, our problem is intuitively well posed, and
222: passes a computational test: we have found no fundamental difficulty
223: in solving models of this type numerically. Furthermore, a careful
224: analysis\cite{torre} of a closely related problem proves that
225: solutions exist and are stable.
226:
227:
228: ``Standing wave'' solutions -- half ingoing and half outgoing --are at
229: the heart of our method, but there is not an obvious definition
230: of standing wave solutions in a nonlinear theory. Our procedure is to
231: find the outgoing ${\cal L}^{-1}_{out}$ and
232: ingoing ${\cal L}^{-1}_{in}$ Green functions for Eq.~(\ref{ourL}).
233: In principle, we can then iterate to find a solution of
234: Eq.~(\ref{ourL}). The iteration
235: \begin{equation}\label{outit}
236: \Psi^{(n+1)}_{out}={\cal L}^{-1}_{out}\big[\sigma(\Psi^{(n)}_{std}
237: )\big]\,,
238: \end{equation}
239: if it converges, gives $\Psi_{out}$,
240: our nonlinear outgoing solution (and similarly for $\Psi_{in}
241: $), while
242: the convergent result of
243: \begin{equation}\label{stdit}
244: \Psi^{(n+1)}_{std}
245: ={\textstyle\frac{1}{2}} \left({\cal L}^{-1}_{out}+{\cal
246: L}^{-1}_{in} \right) \big[\sigma(\Psi^{(n)}_{std} )\big]\,
247: \end{equation}
248: is what we mean by our nonlinear standing wave solution,
249: $\Psi_{std}$. The standing wave solution
250: $\Psi_{std}$
251: is fundamentally different from
252: $(\Psi_{out}+\Psi_{in})/2$, but if effective linearity
253: is correct, the two are very nearly equal.
254: (Note: In practice, for strong nonlinearities, the direct
255: iteration described above must be replaced by Newton-Raphson
256: iteration.)
257:
258: We have previously\cite{paperI} solved the model problem of
259: Eq.~(\ref{ourL}) with more-or-less straightforward finite differencing
260: and direct matrix inversion. (The mixed nature of the partial
261: differential equations prevents the use of such efficient techniques
262: as explicit relaxaton.) This approach was successful (iterations
263: converged) for models with a limited range of source velocities and
264: nonlinearities. More recently we have developed an innovative
265: numerical method that gives remarkably good results, with very little
266: computational cost, and that might be useful in problems other than
267: ours. Our new method is based on three elements. First, we use
268: ``adapted coordinates,'' comoving coordinates that conform to the geometry of
269: the problem. Near the source points our coordinate surfaces approach
270: those of source-centered spherical harmonics; far from the sources the
271: coordinates become spherical polar coordinates centered on the
272: midpoint of the orbit. Our specific choice of adapted coordinates is
273: two-centered bipolar coordinates $\chi,\Theta,\Phi$ (pictured in the
274: equatorial plane of the orbit in Fig.~\ref{fig:2dros}), which
275: asymptotically approach spherical coordinates $r,\theta,\phi$. As
276: discussed in \cite{eigenspec}, this choice is not the most
277: computationally efficient possibility, but it has the advantage of
278: relative simplicity.
279:
280: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
281: \begin{figure}[ht]
282: \includegraphics[width=.26\textwidth]{2dros}
283: \caption{
284: TCBC coordinates in the equatorial plane.
285: \label{fig:2dros}}
286: \end{figure}
287: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
288:
289:
290: The second element of our method is to expand in multipoles (spherical
291: harmonics) of the adapted coordinates. If the separation between the
292: sources is much larger than the size of the source, then the only
293: feature of the source that is important is the ``local'' monopole
294: (monopole in source-centered coordinates). In the wave zone, on the
295: other hand, only the quadrupole of the field matters for the
296: radiation (unless the source motion is highly relativistic). This
297: suggests that a multipole expansion in the adapted coordinates need
298: retain only the monopole and quadrupole. This turns out to be true for
299: source speeds $\leq0.3\,c$. For somewhat larger speeds, good accuracy
300: requires that the hexadecapole be kept in addition. This severe
301: filtering of the multipoles reduces the computational burden of
302: solving the problem, but more important, it eliminates the numerical
303: noise at short angular scales that we found for straightforward finite
304: difference computations.
305:
306: An ``eigenspectral'' treatment of multipoles is the third element of
307: our method that is innovative.
308: A straightforward approach to
309: dealing with multipoles would be to use $Y_{\ell m}(\Theta_i ,\Phi_j
310: )$ on an angular grid of the $\Theta$ and $\Phi $ coordinates, that
311: is, to use the multipoles of the continuum mathematics evaluated on
312: the discrete numerical grid. We have found that this approach does not
313: work at all well for our purposes. Because the monopole is so much
314: greater than the quadrupole in the wave zone, projecting out the
315: quadrupole component with the continuum multipole gives a large error.
316: We have therefore used the multipoles that seem ideally suited to
317: decomposition on the angular grid with $n_\Theta\times n_\Phi$ grid
318: vertices. We view the values of our solution as vectors
319: $\Psi(\Theta_i, \Phi_j)$ in a space of dimension $n_\Theta\times
320: n_\Phi$. At large $\chi$ the angular Laplacian
321: \begin{displaymath}
322: L\equiv \frac{1}{\sin\Theta}\frac{\partial}{\partial\Theta}\left(
323: \sin\Theta\,\frac{\partial}{\partial\Theta}
324: \right)+\frac{1}{\sin^2\Theta}\,\frac{\partial^2
325: }{\partial\Phi^2}
326: \end{displaymath}
327: can be implemented as a finite difference operator, on the
328: $n_\Theta\times n_\Phi$ space, that is self-adjoint with respect to
329: the finite difference equivalent of integration over solid
330: angle\cite{nakamura}. The eigenvectors of this self-adjoint operator
331: are approximately
332: %
333: $Y_{\ell m}(\Theta_i ,\Phi_j)$, but the eigenvectors are exactly
334: orthogonal so that the projection of the radiative multipoles is not
335: affected by the large monopole.
336:
337: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
338: \begin{figure}[ht]
339: \includegraphics[width=.22\textwidth]{fig2a}
340: \includegraphics[width=.22\textwidth]{fig2b} \caption{A comparison of
341: exact (series) linear outgoing solutions with eigenspectral
342: solutions. The radiation field ($\Psi
343: $ minus its monopole) is plotted against the comoving coordinate $\widetilde{Z}
344: $, the distance from the center along a line through the source points.
345: Eigenvectors were found on a $40\times80
346: $ grid for one quadrant, and 16001 radial grid points were used, with an inner
347: boundary condition at
348: $0.02\,a$, and a Sommerfeld outer condition at $80\,a$.
349: \label{fig:compare}}
350: \end{figure}
351: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
352:
353:
354: Questions about the validity of our eigenspectral method apply just as
355: much to the linear version as to the nonlinear version of our model
356: problem, and for the linear problem exact (infinite series) solutions
357: exist for comparison. Figure~\ref{fig:compare} shows such
358: comparisons for
359: models with source velocity $a\Omega=0.4c$ and $a\Omega=0.5c$
360: and shows multipole filtering that allows either the monopole plus
361: quadrupole ($\ell_{max}=3$), or with the hexadecapole included
362: ($\ell_{max}=5$). As the figure indicates, and as should be expected,
363: the number of multipoles that need be included to achieve a given
364: accuracy increases with increasing source speed. The minimal
365: $\ell_{max}=3$ multipole set may be adequate for approximate results
366: at $a\Omega=0.4c$ and gives excellent results for $a\Omega=0.3c$ (not
367: shown). There is, of course, no need to limit the number of multipoles
368: retained to such a small set, but as the number of multipoles
369: approaches the maximum number of eigenvectors that can be found for
370: the angular grid, the smoothing effect of multipole filtering is lost
371: and numerical noise becomes significant. A large number of multipoles
372: would be necessary only for source speeds $a\Omega$ close to $c$. As
373: already mentioned, it is unlikely that such motion will satisfy the
374: slow inspiral condition for which our approximation is designed.
375:
376: Other confirmations of the validity of the method, especially for
377: nonlinear models, have been carried out, and are reported in
378: Ref.~\cite{eigenspec}. We focus here on the most important question
379: that can be answered with these models and numerical methods: Does
380: effective linearity work? Can we extract a good approximation to the
381: outgoing nonlinear problem from the sort of standing wave computation we
382: will be limited to when dealing with Einstein's theory?
383: Figure~\ref{fig:extract} gives strong evidence that we can.
384: For the chosen parameters $\lambda=-15 $ and
385: $\Psi_0=0.15$, nonlinearities are significant, strong enough to reduce field
386: strength by around two-thirds. The outgoing and standing wave
387: solutions were each computed by the Newton-Raphson version of the iteration
388: in Eqs.~(\ref{outit}),(\ref{stdit}).
389:
390: An outgoing approximation is extracted from the standing wave solution
391: by the following steps: (i)~In the outer region the solution is
392: matched to a general linear standing wave solution of half-ingoing and
393: half-outgoing waves; an outgoing wave is extracted as if the problem
394: were linear. (ii)~In the strong-field region very close to the source,
395: the standing wave solution itself is taken to be the outgoing
396: approximation. (iii)~The two solutions are blended over a narrow
397: intermediate range of radii.
398:
399: In Fig.~\ref{fig:extract}, the computed outgoing nonlinear solution is
400: shown as a solid curve. The data-type points representing the outgoing
401: wave extracted from the standing wave solution show how good the
402: approximation is. We have run models with much stronger nonlinearity
403: and have found equally good, or better, agreement of the true outgoing
404: solution and the extracted approximation. The validity of effective
405: linearity should, in fact, become questionable not for stronger
406: nonlinearity, but only for physically implausible high source velocity.
407:
408: In addition to confirming effective linearity, computation with the
409: model has also allowed some early insights about sensitivity to
410: source details. By varying the multipole content of the inner boundary data
411: we explored the impact of source structure on the radiation
412: field. The result (detailed in Ref.~\cite{eigenspec}) is in perfect
413: accord with physical intuition; the radiation is insensitive to
414: source structure unless the source size becomes comparable to the
415: separation of the sources (i.e.\,, unless the moments ascribable to
416: the structure of the individual sources are comparable to the
417: quadrupole moment due to the separation of the mass points). The
418: equivalent question for Einstein's theory is more difficult, but we
419: should be able to give clear quantitative answers.
420:
421: The next steps in our program start with linearized gravity in the harmonic
422: gauge. We have already done this with a finite difference code; work
423: on applying the eigenspectral method is underway, and no significant
424: problems are anticipated. The infrastructure of the linearized gravity
425: will provide much of what is needed for post-Minkowskian computations,
426: since the structure of the linear operators (the analogs of the ${\cal
427: L} $s in Eqs.~(\ref{ourL})--(\ref{stdit})) will be the same as for
428: linearized gravity. The final step to the full Einstein theory will,
429: in the same way, be based on the numerical infrastructure developed
430: for post-Minkowsian models.
431:
432:
433:
434:
435:
436:
437:
438:
439: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
440: \begin{figure}[ht]
441: \includegraphics[width=.23\textwidth]{fig3a}
442: \includegraphics[width=.23\textwidth]{fig3b} \caption{
443: The computed nonlinear outgoing solution compared with an approximate
444: outgoing solution extracted from the computed nonlinear standing wave
445: solutions. Grid parameters are the same as for Fig.~\ref{fig:compare},
446: and $\ell_{\rm max}
447: =5
448: $ was used.
449: \label{fig:extract}}
450: \end{figure}
451: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
452:
453:
454: We gratefully acknowledge the support of NSF grant PHY0244605 and NASA
455: grant ATP03-0001-0027, to UTB and of NSF grant PHY-0099568 and NASA
456: grant NAG5-12834 to Caltech.
457: %
458: We have greatly benefited from suggestions with John Friedman, Christopher
459: Beetle, and Lior Burko.
460:
461:
462:
463: \begin{thebibliography}{10}
464: \bibitem{laz1}
465: J. Baker {\it et~al.}, Phys. Rev. D {\bf 65}, 124012 (2002).
466:
467: \bibitem{PN} L.~Blanchet,
468: Living Rev. Relativity 5, (2002), 3. [Online article]: cited Dec.~25,
469: 2000 http://www.livingreviews.org/lrr-2002-3.
470:
471:
472:
473: \bibitem{det}
474: J.~K. Blackburn and S. Detweiler, Phys. Rev. D {\bf 46}, 2318 (1992).
475:
476:
477: \bibitem{det2}
478: S. Detweiler, Phys. Rev. D {\bf 50}, 4929 (1994).
479:
480:
481: \bibitem{WKP}
482: J.~T. Whelan, W. Krivan, and R.~H. Price, Class. Quant. Grav. {\bf 17}, 4895
483: (2000).
484:
485: \bibitem{WBLandP}
486: J.~T. Whelan, C. Beetle, W. Landry, and R.~H. Price, Class. Quant. Grav. {\bf
487: 19}, 1285 (2002).
488:
489: \bibitem{rightapprox}
490: R.~H. Price, Class. Quant. Grav. {\bf 21}, S281 (2004).
491:
492:
493: \bibitem{paperI}
494: Z. Andrade {\it et~al.}, Phys. Rev. D {\bf 70}, 064001 (2004).
495:
496:
497: \bibitem{friedmanetal}
498: M. Shibata, K. Uryu, J. L. Friedman, Phys. Rev. D {\bf 70}, 044044 (2004).
499: Erratum-ibid. D {\bf70} 129901
500: (2004).
501:
502: \bibitem{torre}
503: C.~G.~Torre, J.~Math.~Phys., {\bf44}
504: 6223-6232 (2003).
505:
506: \bibitem{eigenspec}
507: B. Bromley, R. Owen and R.~H.~Price, submitted
508: to Phys.~Rev.~D. Preprint gr-qc/0502034.
509:
510: \bibitem{nakamura}
511: After completing our scalar work with the eigenspectral method, we
512: discovered that essentially the same technique was used by
513: T. {Nakamura}, Progress of Theoretical Physics {\bf 72}, 746 (1984).
514:
515: \end{thebibliography}
516:
517:
518: \end{document}
519:
520:
521:
522:
523:
524: