1: %=========================================================================
2: \section{Numerical evolution of excited cosmic strings}
3: \label{cstr}
4: %
5: %
6: %==================================
7: \subsection{Introduction}
8: %
9: %
10: According to the standard ``big bang'' model of cosmology, the universe
11: is continuously expanding and cooling and was extremely hot and dense in its
12: early stages. The grand unified theories (GUT) of elementary particle
13: physics predict phase transitions to occur as a result of this cooling
14: process in the early universe. These result in topological defects, regions
15: with the ``old symmetry'' surrounded by ``new symmetry''. The topology
16: of the defects depends on the symmetry groups characterising the
17: involved fields before and after the symmetry breaking. Cosmic strings
18: are a 1-dimensional, ``string-like'' version of these topological defects.
19: The type of strings usually considered from the astrophysical point of view
20: has a mass per unit length $\mu \approx 10^{-6}$ in natural units
21: ($\hbar = G = c = 1$). The corresponding phase transitions are predicted
22: to have occurred at the GUT energy scale $10^{15}$ GeV. Strings with
23: significantly higher mass created at higher energy scales cannot be ruled
24: out, however, and their treatment can no longer be achieved in the
25: weak-field limit. \\
26: Numerical simulations by \scite{Vachaspati1984} show that
27: cosmic strings are created in the form of a network of infinitely
28: long or loop like strings. In this work we will focus on infinitely
29: long strings which are modelled in the framework of cylindrical symmetry. \\
30: Cosmic strings have caught the interest of astrophysicists and
31: relativists for several reasons. Most importantly the suggestion
32: that cosmic strings be seeds for galaxy formation by \scite{Zeldovich1980}
33: has given rise to intense efforts to understand the evolution of the
34: resulting density perturbations (see e.g. \citeNP{Turok1986}).
35: Cosmic strings are also thought to be sources of gravitational radiation
36: (\citeNP{Vilenkin1994}). Below we will study the interaction of an
37: infinitely long cosmic string with a wave pulse with one gravitational
38: degree of freedom.
39: Cosmic strings have also been considered of astrophysical relevance
40: because of the bending of light rays that arises from the conical structure
41: of the resulting spacetime.
42: It has been shown by \scite{Vilenkin1981} that the
43: geometry around an isolated cosmic string is Minkowskian minus
44: a wedge, the ``deficit angle'', and consequently cosmic strings
45: may act as gravitational lenses. \\
46: Even though static cosmic strings in cylindrical symmetry have been
47: studied extensively in the past either in Minkowskian or curved
48: spacetime (see e.g. \shortciteNP{Laguna1987},
49: \citeNP{Garfinkle1985}), no solution
50: has been obtained, to our knowledge, for a dynamic cosmic string
51: coupled to gravity via the fully non-linear Einstein equations. Below we
52: will present a numerical solution of this scenario and investigate
53: the behaviour of a cosmic string excited by gravitational radiation.
54: After presenting the mathematical description of a cosmic string
55: in the next section we will derive the equations of a dynamic
56: cosmic string coupled to gravity. In section \ref{CSnum} we will
57: describe the numerical treatment of these equations.
58: The simple scenario of a static cosmic string in Minkowski spacetime
59: presents already most of the subtleties involved
60: in solving the general problem and is therefore suitable for illustrating our
61: numerical methods. Subsequently we address a static string in curved
62: spacetime and finally present the dynamic code. This code is
63: extensively tested in section \ref{CStest} before we investigate
64: the time evolution in section \ref{CSevol}. \\
65: The results and techniques presented in this section can also be found in
66: \scite{Sperhake2000}. \\
67: We conclude this introduction with some
68: comments on the numerical formulations used in this section. We have seen
69: above how the combination of an interior Cauchy evolution with a
70: characteristic evolution in the exterior region leads to a stable
71: accurate simulation of cylindrically symmetry vacuum spacetimes. In a
72: natural extension of this project we studied the inclusion of matter in
73: the form of a cylindrically symmetric cosmic string. Such an
74: extension of the CCM-code of the previous section has been
75: developed, but no long term stable evolutions have been achieved
76: with that code. Consequently we have restarted the investigation.
77: For convenience this has been done in a purely characteristic
78: framework and finally
79: resulted in the long-term stable, accurate code described below. In the
80: course of this work we have isolated the existence of exponentially
81: diverging solutions and the corresponding difficulties at the outer boundary
82: as the source of the problems. We will describe how these difficulties
83: can be naturally controlled with the use of implicit numerical techniques.
84: The use of such techniques, however, is by no means restricted to
85: characteristic methods and we have no reason to believe that an implicit
86: Cauchy-characteristic matching code would perform less satisfactorily.
87: Such an implicit CCM code has been tested in the simple case
88: of a cylindrically symmetric vacuum spacetime with vanishing rotation
89: and has lead to an accurate long-term stable evolution of the
90: Weber-Wheeler wave. From this point of view the choice of a characteristic
91: formulation for the work described in this section is merely a
92: consequence of the chronology in which progress has been achieved. \\
93:
94:
95: %=========================================================================
96: \subsection{Mathematical description of a cosmic string}
97: %
98: %
99: In the following work we will use cylindrical coordinates $r$, $\phi$,
100: $z$. Here $z$ is the Killing direction corresponding to cylindrical
101: symmetry and $r$, $\phi$ are standard polar coordinates. In 4-dimensional
102: spacetime the time coordinate is $t$, but we will apply a characteristic
103: formalism for the numerical solution and therefore also use the retarded
104: time $u=t-r$.
105: The simplest model of a cosmic string consists of a scalar field
106: $\Phi$ coupled to a $U(1)$-gauge field $\hbox{\vec A}_{\mu}$.
107: The Lagrangian for these coupled fields is given by
108: %
109: \begin{align}
110: L_M &= -|(\nabla_{\mu}+ie\hbox{\vec A}_{\mu})\Phi|^2 - V(\Phi)
111: - \frac{1}{4} \hbox{\vec F}_{\mu \nu}\hbox{\vec F}^{\mu \nu}
112: \label{LM}.
113: \end{align}
114: %
115: Here $e$ is a constant, which describes the coupling between the
116: scalar and the vector field.
117: The self-coupling potential $V(\Phi)$ has the ``Mexican-hat'' shape
118: predicted by the standard model of elementary particle physics and
119: $\hbox{\vec F}_{\mu \nu}$ is the field tensor
120: %
121: \begin{align}
122: \hbox{\vec F}_{\mu \nu} &= \nabla_{\mu}\hbox{\vec A}_{\nu}
123: - \nabla_{\nu}\hbox{\vec A}_{\mu}, \\[10pt]
124: V(\Phi) &= 2 \lambda (\Phi^2 - \langle \Phi \rangle^2)^2,
125: \end{align}
126: %
127: where $\lambda$ is the self-coupling constant of the scalar field.
128: It turns out to be useful to introduce the Higgs vacuum expectation
129: value of the scalar field as a parameter $\eta = 2 \langle
130: \Phi \rangle^2$.
131: Generalizing the notation of \scite{Garfinkle1985} we write the fields as
132: %
133: \begin{align}
134: \Phi &= \frac{S}{\sqrt{2}} e^{i\psi}, \label{Phi} \\[10pt]
135: \hbox{\vec A}_{\mu} &= \frac{1}{e} (P-1) \nabla_{\mu} \phi , \label{A}
136: \end{align}
137: %
138: where $P$, $S$ and $\psi$ are functions of $u$, $r$, $\phi$.
139: From now on, however, we will make the simplifying assumption of
140: cylindrical symmetry. Then $P$ and $S$ are functions of
141: $u$, $r$ only and $\psi = n\phi$, where $n$ is the winding number. In this
142: work we will only consider the case $n=1$, so $\psi = \phi$.
143: We can calculate the energy momentum tensor $\hbox{\vec T}^{\mu \nu}$
144: from the Lagrangian according to
145: %
146: \begin{align}
147: \hbox{\vec T}^{\mu \nu} &= \frac{2}{\sqrt{-\hbox{\vec g}}}
148: \frac{\delta \mathcal{L}_M} {\delta \hbox{\vec g}_{\mu \nu}},
149: \end{align}
150: %
151: where $\mathcal{L}_M = \sqrt{-\hbox{\vec g}} L_M$ is the Lagrange density.
152: Summarising the variables and parameters, we have
153: %
154: \begin{list}{\rm{(\arabic{count})}}{\usecounter{count}}{
155: \labelwidth1cm \leftmargin1.5cm \labelsep0.4cm \rightmargin1cm
156: \parsep0.5ex plus0.2ex minus0.1ex \itemsep0ex plus0.2ex}
157: \item the amplitude of the scalar field $S(u,r)$,
158: \item the amplitude of the $U(1)$ gauge field $P(u,r)$,
159: \item the constant $e$ which describes the coupling between
160: the scalar and vector field,
161: \item the self-coupling constant $\lambda$ of the scalar field,
162: \item the vacuum expectation value of the scalar field $\eta$.
163: \end{list}
164: %
165: If we substitute Eqs.\,(\ref{Phi}), (\ref{A}) in (\ref{LM}) we obtain the
166: Lagrangian and the energy momentum tensor in terms of these quantities
167: %
168: \begin{align}
169: L_M &= -\frac{1}{2} \hbox{\vec g}^{\mu \nu} (\nabla_{\mu} S) (\nabla_{\nu} S)
170: -\frac{1}{2} S^2 \hbox{\vec g}^{\mu \nu} (\nabla_{\mu} \phi
171: + e\hbox{\vec A}_{\mu})
172: (\nabla_{\nu} \phi +e\hbox{\vec A}_{\nu}) - \lambda (S^2-\eta^2)^2
173: -\frac{1}{4} \hbox{\vec F}_{\mu \nu} \hbox{\vec F}^{\mu \nu}, \\[10pt]
174: \hbox{\vec T}_{\mu \nu} &= (\nabla_{\mu}S) (\nabla_{\nu}S)
175: + S^2(\nabla_{\mu} \phi + e\hbox{\vec A}_{\mu})(\nabla_{\nu} \phi
176: +e\hbox{\vec A}_{\nu})
177: +\hbox{\vec g}_{\mu \nu} L_M \label{CSemtensor} .
178: \end{align}
179: %
180:
181: %=========================================================================
182: \subsection{The field equations}
183: %
184: %
185: We start again with the line element in Jordan, Ehlers, Kundt
186: and Kompaneets (JEKK) form (\ref{JEKK}) for a cylindrically
187: symmetric spacetime. This form of the metric, however, is not compatible
188: with the cosmic string energy momentum tensor so we follow
189: \scite{Marder1958} by introducing an extra variable $\mu$ into the metric
190: %
191: \begin{equation}
192: ds^2 = e^{2(\gamma-\psi)}(-d\tilde{t}^2 + d\tilde{r}^2)
193: + \tilde{r}^2e^{-2\psi}d\phi^2
194: + e^{2(\psi+\mu)}(\omega d\phi+dz)^2,
195: \end{equation}
196: %
197: where the tilde is used to reserve the names
198: $t$ and $r$ for rescaled coordinates below.
199: This choice enables us to compare our numerical solutions
200: with the results of the Cauchy-characteristic matching code
201: described in section \ref{ccm}. We have already
202: noted that this metric has a zero shift vector and the lapse is determined
203: by the requirement $\hbox{\vec g}_{\tilde{t}\tilde{t}}=
204: \hbox{\vec g}_{\tilde{r}\tilde{r}}$.
205: The function $\mu$, however, introduces
206: the extra gauge freedom of relabelling the radial null surfaces:
207: $\tilde{u} \rightarrow f(\tilde{u})$ and $\tilde{v}
208: \rightarrow g(\tilde{v})$. We may fix this
209: by specifying the initial values for $\mu$ and either its time derivative
210: in a ``3+1'' formalism or its boundary conditions in a characteristic
211: formalism. We will follow the second approach and below we will see that
212: the function $\mu$ is uniquely determined in the static case and
213: the boundary conditions follow from regularity assumptions
214: of the metric. The further requirement that the dynamic results reduce
215: to the static ones in the case of vanishing time dependence therefore
216: fixes the gauge. \\
217: It turns out that we can eliminate one of the free parameters and simplify
218: the equations if we introduce rescaled quantities according to
219: %
220: \begin{align}
221: t &= \sqrt{\lambda}\eta \t, \label{rescale_t} \\[10pt]
222: r &= \sqrt{\lambda}\eta \r, \\[10pt]
223: X &= \frac{S}{\eta}, \label{rescale_X}\\[10pt]
224: \alpha &= \frac{e^2}{\lambda}.
225: \end{align}
226: %
227: Thus $\alpha$ represents the relative strength of the coupling between
228: scalar and vector field compared to the self-coupling. Furthermore
229: we use the retarded time $u = t - r$ so that the line element becomes
230: %
231: \begin{align}
232: ds^2 &= \frac{e^{2(\gamma-\psi)}}{\lambda \eta^2}
233: (-du^2 - 2du dr) + r^2\frac{e^{-2\psi}}
234: {\lambda \eta^2}d\phi^2 + e^{2(\psi+\mu)}(\omega d\phi+dz)^2.
235: \end{align}
236: %
237: In section \ref{Gerochdec} we have described the Geroch decomposition which
238: can be used to factor out the Killing direction $\bbox{\partial}_z$ even
239: if the Killing field is not hypersurface-orthogonal. It is a remarkable
240: fact that the right hand side of equation (\ref{Geroch_cond}) still
241: vanishes for spacetimes with a cosmic string energy-momentum tensor
242: (\ref{CSemtensor}) (\shortciteNP{Sjodin2000}),
243: so that the Geroch twist can be described by a
244: potential according to Eq.\,(\ref{Geroch_pot}). The other geometrical variable,
245: the norm of the $z$-Killing vector (\ref{Geroch_norm}) becomes
246: %
247: \begin{align}
248: \nu = e^{2(\psi + \mu)},
249: \end{align}
250: %
251: and the 3-dimensional line element (\ref{Geroch_3m}) is
252: %
253: \begin{align}
254: ds^2 &= \frac{1}{\lambda \eta^2 \nu} \left[e^{2(\gamma-\psi)}
255: (-du^2 - 2du dr) + r^2e^{2\mu}
256: d\phi^2 \right] \label{CS_3dline}.
257: \end{align}
258: %
259: With the energy momentum tensor given by (\ref{CSemtensor}) and the
260: 3-dimensional line element (\ref{CS_3dline}) we are now in a position to
261: calculate the field equations according to equations
262: (\ref{Geroch3d})-(\ref{Geroch_tau}). We obtain
263: %
264: %
265: \begin{align}
266: \begin{split}
267: \Box \nu =& \,\, \nu_{,r} \mu_{,r} + \frac{\tau_{,r}^2-\nu_{,r}^2}{\nu}
268: -\nu_{,u} \mu_{,r} - \nu_{,r} \mu_{,u}
269: +2\frac{\nu_{,u} \nu_{,r} - \tau_{,u} \tau_{,r}}{\nu} \\[10pt]
270: & +8\pi \eta^2\left[ 2e^{2(\gamma+\mu)} (X^2-1)^2
271: +e^{-2\mu} \nu^2
272: \frac{2P_{,u} P_{,r} - P_{,r}^2}{\alpha r^2} \right], \label{nuur}
273: \end{split} \\[10pt]
274: \Box \tau =&\,\, \tau_{,r}\mu_{,r} - 2\frac{\tau_{,r} \nu_{,r}}{\nu}
275: - \tau_{,u} \mu_{,r}-\tau_{,r} \mu_{,u}
276: + 2\frac{\tau_{,r} \nu_{,u} +\tau_{,u} \nu_{,r}}{\nu}, \\[10pt]
277: \Box \mu =&\,\, \mu_{,r}^2 +\frac{\mu_{,r}}{r}
278: - \frac{\mu_{,u}}{r} -2\mu_{,u} \mu_{,r} +8\pi\eta^2 \left[
279: 2\frac{e^{2(\gamma + \mu)}}{\nu} (X^2-1)^2 + e^{2\gamma}
280: \frac{X^2P^2}{r^2} \right] \\[10pt]
281: 0 =&\,\, 2\gamma_{,r} +2r\gamma_{,r} \mu_{,r} -r\mu_{,rr} + r\mu_{,r}^2
282: -\frac{r}{2\nu^2} (\tau_{,r}^2 + \nu_{,r}^2) - 8\pi \eta^2 \left[
283: r X_{,r}^2 + \frac{1}{\alpha} e^{-2\mu} \nu \frac{P_{,r}^2}{r} \right]
284: \label{gammar},
285: \end{align}
286: %
287: where we have introduced the flat-space d'Alembert operator
288: %
289: \begin{equation}
290: \Box = 2\frac{\partial^2}{\partial u \partial r}
291: - \frac{\partial^2}{\partial r^2} - \frac{1}{r} \left(
292: \frac{\partial}{\partial r} - \frac{\partial}{\partial u} \right).
293: \end{equation}
294: %
295: This set of equations is supplemented by the matter evolution equations
296: obtained either from conservation of energy-momentum
297: $\nabla_{\mu} \hbox{\vec T}^{\mu \nu} = 0$ or variation of the Lagrange density
298: $\mathcal{L}_M$ with respect to the matter fields $P$ and $X$. The result is
299: %
300: \begin{align}
301: \Box P &= 2\frac{P_{,u}-P_{,r}}{r} - P_{,r} \mu_{,r}
302: + P_{,r}\frac{\nu_{,r}}{\nu} + P_{,r} \mu_{,u} + P_{,u} \mu_{,r}
303: - \frac{P_{,r} \nu_{,u} + P_{,u} \nu_{,r}}{\nu}
304: - \alpha \frac{e^{2(\gamma + \mu)}}{\nu} PX^2,
305: \label{pur} \\[10pt]
306: \Box X &= X_{,r} \mu_{,r} - X_{,u} \mu_{,r} - X_{,r} \mu_{,u}
307: - 4 \nu^{-1} e^{2(\gamma + \mu)} X(X^2-1)
308: - e^{2\gamma} \frac{XP^2}{r^2}. \label{xur}
309: \end{align}
310: %
311: Note that in equations (\ref{nuur})-(\ref{gammar}) the matter terms
312: exclusively appear with a factor $\eta^2$. Consequently $\eta$ describes
313: the effect of the string on the spacetime geometry and, thus, represents the
314: string's mass.
315: There are two further Einstein equations which can be shown to be
316: a direct consequence of (\ref{nuur})-(\ref{xur}) and their derivatives.
317: These equations have only been used to provide a check on the accuracy
318: of the code.
319: Finally we have to supplement the equations by
320: boundary conditions on the axis. For the 4-dimensional metric
321: variables the simplest condition is to require the metric to be
322: $C^2$ on the axis so that we have a well defined curvature tensor.
323: The resulting boundary conditions are (\citeNP{Sjodin2001b})
324: %
325: \begin{align}
326: \nu(t,r) &= a_1(t) + \mathscr{O}(r^2), \label{bound_nutr} \\[10pt]
327: \tau(t,r) &= \mathscr{O}(r^2), \\[10pt]
328: \mu(t,r) &= a_2(t) + \mathscr{O}(r^2), \label{bound_mutr} \\[10pt]
329: \gamma(t,r) &= \mathscr{O}(r). \label{gammatr}
330: \end{align}
331: %
332: The boundary conditions for $S$ and $P$ are (\citeNP{Garfinkle1985})
333: %
334: \begin{align}
335: P(t,r) &= 1+\mathscr{O}(r^2), \\[10pt]
336: X(t,r) &= \mathscr{O}(r). \label{bound_xtr}
337: \end{align}
338: %
339: The numerical implementation of these boundary conditions as well as regularity
340: requirements at null infinity will be discussed in section \ref{dynstring}.
341:
342:
343: %=======================================================================
344: \subsection{Numerical methods}
345: \label{CSnum}
346: %
347: %
348: In order to solve the above field equations we have developed two
349: independent codes. The first is based on
350: the Cauchy characteristic matching code described in section \ref{ccm}.
351: This code performs well in the absence of matter and has been used
352: to study several cylindrically symmetric vacuum solutions
353: (see also \shortciteNP{Sjodin2000}). However, this CCM code performed less
354: satisfactorily in the evolution of the cosmic string. This
355: is due to the existence of
356: unphysical solutions to the evolution equations (\ref{nuur})-(\ref{xur}) which
357: diverge exponentially as $r\rightarrow \infty$. Controlling the time evolution
358: near null infinity by means of a sponge function enabled us to select the
359: physical solutions with regular behaviour at $I^+$, but the sponge function
360: itself introduced noise which eventually gave rise to instabilities.
361: We therefore implemented a second implicit,
362: purely characteristic, code which
363: allows us to directly control the behaviour of the solutions at the
364: boundaries and thus suppress diverging solutions. The main problem with
365: the system of differential equations is the irregularity of the equations
366: at both the origin and null infinity. It is the implicit nature of the
367: scheme that provides a
368: simple way of implementing the boundary conditions and thus circumventing
369: all problems with these irregularities. A purely characteristic formulation
370: has been used for the second code for convenience rather than
371: numerical necessity and we believe that an implicit CCM scheme would
372: produce similar accuracy, convergence and long term stability.
373: It is interesting that the irregularity problems are
374: already present in the calculation of the static
375: cosmic string in Minkowski spacetime. We will,
376: therefore, first describe the numerical scheme used in the static Minkowskian
377: case where the equations are fairly simple. We then present the modifications
378: necessary for the static and dynamic case coupled to the gravitational
379: field.
380:
381:
382: %========================================================================
383: \subsubsection{The static cosmic string in Minkowski spacetime}
384: \label{SECcsmink}
385: %
386: %
387: In Eqs.\,(\ref{nuur})-(\ref{xur})
388: we set the metric variables to their Minkowskian values and all time
389: derivatives to zero to obtain the equations for the static cosmic string in
390: Minkowski spacetime (cf.\ \citeNP{Garfinkle1985})
391: %
392: \begin{align}
393: r\frac{d}{dr} \left( r^{-1} \frac{dP}{dr} \right) &= \alpha X^2 P
394: \label{mink_Prr}, \\[10pt]
395: r\frac{d}{dr}\left( r\frac{dX}{dr} \right) &= X \left[ P^2 + 4r^2 (X^2-1)
396: \right] . \label{mink_Xrr}
397: \end{align}
398: %
399: The boundary conditions are (see \citeNP{Garfinkle1985})
400: %
401: \begin{alignat}{4}
402: P(0) &= 1, & \qquad \qquad & \lim_{r\rightarrow \infty} P(r) &= 0,
403: \nonumber \\
404: X(0) &= 0, & &\lim_{r \rightarrow \infty} X(r) &= 1.
405: \label{mink_bound}
406: \end{alignat}
407: %
408: In order to cover the whole spacetime with a finite coordinate range,
409: we divide the computational domain into two regions in the same way
410: as in section \ref{start_int}. In the inner region ($0\le r\le 1$) we
411: use the coordinate $r$, while in the outer region we introduce the compactified
412: radius $y$ defined by equation (\ref{y})
413: which covers the range $1 \ge y \ge 0$. This corresponds to the region
414: $1 \le r < \infty$ with infinity mapped to $y=0$.
415: Again we combine $r$ and $y$ into the
416: single radial variable $w$ defined by (\ref{w}).
417: In terms of the coordinate $y$ Eqs.\,(\ref{mink_Prr}),
418: (\ref{mink_Xrr}) take the form
419: %
420: \begin{eqnarray}
421: &&\frac{d}{dy} \left( y^5\frac{dP}{dy} \right) = 4\alpha \frac{X^2P}{y},
422: \\[10pt]
423: &&\frac{d}{dy} \left( y \frac{dX}{dy} \right) =
424: 4X\left[\frac{P^2}{y}+ 4\frac{(X^2-1)}{y^5}\right].
425: \end{eqnarray}
426: %
427: The number of grid points in each region may differ, but
428: %
429: \begin{figure}[t]
430: \centering
431: %------------------------------------------------------------------------
432: \begin{picture}(0,0)%
433: \epsfig{file=grid2.ps}%
434: \end{picture}%
435: \setlength{\unitlength}{3947sp}%
436: %
437: \begingroup\makeatletter\ifx\SetFigFont\undefined%
438: \gdef\SetFigFont#1#2#3#4#5{%
439: \reset@font\fontsize{#1}{#2pt}%
440: \fontfamily{#3}\fontseries{#4}\fontshape{#5}%
441: \selectfont}%
442: \fi\endgroup%
443: \begin{picture}(6312,2028)(976,-1606)
444: \put(3226,-436){\makebox(0,0)[lb]{\smash{\SetFigFont{11}{13.2}{\rmdefault}{\mddefault}{\updefault}...}}}
445: \put(2401,-436){\makebox(0,0)[lb]{\smash{\SetFigFont{11}{13.2}{\rmdefault}{\mddefault}{\updefault}3}}}
446: \put(1801,-436){\makebox(0,0)[lb]{\smash{\SetFigFont{11}{13.2}{\rmdefault}{\mddefault}{\updefault}2}}}
447: \put(976,-436){\makebox(0,0)[lb]{\smash{\SetFigFont{11}{13.2}{\rmdefault}{\mddefault}{\updefault}$k$=1}}}
448: \put(976,-61){\makebox(0,0)[lb]{\smash{\SetFigFont{11}{13.2}{\rmdefault}{\mddefault}{\updefault}$r$=0}}}
449: \put(3976,-1186){\makebox(0,0)[lb]{\smash{\SetFigFont{11}{13.2}{\rmdefault}{\mddefault}{\updefault}$K_1$+1}}}
450: \put(4051,-436){\makebox(0,0)[lb]{\smash{\SetFigFont{11}{13.2}{\rmdefault}{\mddefault}{\updefault}$K_1$}}}
451: \put(4051,-61){\makebox(0,0)[lb]{\smash{\SetFigFont{11}{13.2}{\rmdefault}{\mddefault}{\updefault}$r$=1}}}
452: \put(4051,-1561){\makebox(0,0)[lb]{\smash{\SetFigFont{11}{13.2}{\rmdefault}{\mddefault}{\updefault}$y$=1}}}
453: \put(4651,-1186){\makebox(0,0)[lb]{\smash{\SetFigFont{11}{13.2}{\rmdefault}{\mddefault}{\updefault}$K_1$+2}}}
454: \put(5251,-1186){\makebox(0,0)[lb]{\smash{\SetFigFont{11}{13.2}{\rmdefault}{\mddefault}{\updefault}$K_1$+3}}}
455: \put(5701,314){\makebox(0,0)[lb]{\smash{\SetFigFont{11}{13.2}{\rmdefault}{\mddefault}{\updefault}outer region}}}
456: \put(2101,314){\makebox(0,0)[lb]{\smash{\SetFigFont{11}{13.2}{\rmdefault}{\mddefault}{\updefault}inner region}}}
457: \put(6901,-1186){\makebox(0,0)[lb]{\smash{\SetFigFont{11}{13.2}{\rmdefault}{\mddefault}{\updefault}$K_1$+$K_2$}}}
458: \put(7051,-1561){\makebox(0,0)[lb]{\smash{\SetFigFont{11}{13.2}{\rmdefault}{\mddefault}{\updefault}$y$=0}}}
459: \put(6376,-1186){\makebox(0,0)[lb]{\smash{\SetFigFont{11}{13.2}{\rmdefault}{\mddefault}{\updefault}...}}}
460: \end{picture}
461:
462: \vspace{0.5cm}
463: %------------------------------------------------------------------------
464: \caption{The combined grid of the inner and the outer region. Note that
465: both grid points, $K_1$ and $K_1+1$, correspond to the position
466: $r=1 \Leftrightarrow y=1$.
467: These points form the interface of the code and facilitate
468: transformation of the variables from the coordinate system using $r$
469: into that using $y$.}
470: \label{grid}
471: \end{figure}
472: %
473: each half-grid is uniform. Thus we use
474: a total of $K:=K_1 + K_2$ grid points where the points labelled $K_1$ and
475: $K_1+1$ both correspond to the position $r=1=y$.
476: The points $K_1$, $K_1+1$ form the interface between the two regions
477: (see Fig.\,\ref{grid}). One point will contain
478: the variables in terms of $r$, the other in terms of $y$.
479: With the computational grid covering the whole spacetime, we now face a
480: two point boundary value problem. Due to the existence of unphysical solutions
481: diverging at $y=0$
482: % shooting methods turned out to be unsuitable for solving
483: %this problem. On the other hand numerical relaxation,
484: we have chosen to solve the equations with a numerical relaxation scheme
485: as described in section \ref{relaxation} which
486: allows us to directly control the behaviour of $P$ and $X$ at infinity.
487: The form of Eqs.\,(\ref{mink_Prr}),\ (\ref{mink_Xrr}) suggests that
488: in order to write them as a first order system we should introduce the
489: auxiliary variables $Q=r^{-1}P_{,r}$ and $R=rX_{,r}$. The equations may then be
490: written in the form
491: %
492: \begin{align}
493: P_{,r} &= rQ, \\[10pt]
494: X_{,r} &= \frac{R}{r}, \\[10pt]
495: Q_{,r} &= \alpha \frac{PX^2}{r}, \\[10pt]
496: R_{,r} &= X\left[\frac{P^2}{r} + 4r(X^2-1)\right].
497: \end{align}
498: %
499: The corresponding equations in the outer region are given by
500: %
501: \begin{align}
502: P_{,y} &= -2\frac{Q}{y^5}, \\[10pt]
503: X_{,y} &= -2\frac{R}{y}, \\[10pt]
504: Q_{,y} &= -2 \alpha \frac{X^2P}{y}, \\[10pt]
505: R_{,y} &= -2X \left( \frac{P^2}{y}+4\frac{X^2-1}{y^5} \right).
506: \end{align}
507: %
508: Standard second order centred finite differencing
509: according to Eqs.\,(\ref{relax_FDE1}), (\ref{relax_FDE2})
510: % Note, however, that in this case
511: % we obtain $4(N-2)$ algebraic equations
512: % (instead of $4(N-1)$ as in section \ref{relaxation})
513: % due to the presence of the interface.
514: results in $4(K-2)$ non-linear algebraic equations which are supplemented
515: by the 4 boundary conditions
516: (\ref{mink_bound}) and 4 interface relations
517: %
518: \begin{align}
519: P_{K_1+1} &= P_{K_1}, \label{int_P} \\[10pt]
520: X_{K_1+1} &= X_{K_1}, \\[10pt]
521: Q_{K_1+1} &= Q_{K_1}, \\[10pt]
522: R_{K_1+1} &= R_{K_1} \label{int_R}.
523: \end{align}
524: %
525: We then start with piecewise linear initial guesses for $P$ and $X$ (and the
526: corresponding derivatives $Q$ and $R$) and solve the $4K$ algebraic equations
527: as described in section \ref{relaxation}. \\
528: In order to check the code for convergence, we vary the grid resolution
529: $K$ (using $K_1=K_2$ points in both regions) from
530: $150$ to $2400$, halving the grid spacing each time.
531: Since we do not have an analytic solution, the results are
532: compared against the high-resolution case ($K=2400$). For doing this we
533: calculate the $\ell_2$ norm according to Eq.\,(\ref{l2norm}).
534: In this case the function $\Psi$ in Eq.\,(\ref{l2norm})
535: stands for $P$, $X$, $Q$ or $R$ and the norm does not depend on time
536: because of the static nature of the problem.
537: For second order convergence we expect the $\ell_2$ norm to decrease by
538: a factor of 4 each time we increase the grid resolution by a factor of 2.
539: However, we do not compare our results against the exact solution but against
540: a high resolution result which itself has a finite truncation error,
541: so that
542: %
543: \begin{align}
544: \ell_2[\Psi^K] &= \left(\sum_k\Psi^K_k - \Psi^{2400}_k\right)^{1/2}.
545: \label{CAUCHYNORM}
546: \end{align}
547: %
548: Therefore we do not expect the factor to be exactly $4$. Using a
549: %
550: \begin{table}[t]
551: \caption{Convergence test for the cosmic string in Minkowski space-time
552: for $\alpha = 1$. The norm of the deviation
553: $\ell_2[\Delta \Psi^K]$ is defined by Eq.\,(\ref{CAUCHYNORM}).
554: As the grid resolution is increased, the deviation from the
555: high resolution result decreases quadratically to a good
556: approximation (see text for details).}
557: \begin{center}
558: \label{convMink}
559: \begin{tabular}{l|cccc}
560: \hline \hline
561: & P & X & Q & R \\
562: \hline\\[-7pt]
563: $\ell_2(\Delta \Psi^{1200})$ & $5.77\cdot 10^{-7}$ & $2.84\cdot 10^{-7}$
564: & $5.86\cdot 10^{-7}$ & $8.89\cdot 10^{-7}$ \\
565: \hline\\[-7pt]
566: $\ell_2(\Delta\Psi^{150})/\ell_2(\Delta\Psi^{300})$
567: & 4.05 & 4.05 & 4.04 & 4.05 \\
568: $\ell_2(\Delta\Psi^{300})/\ell_2(\Delta\Psi^{600})$
569: & 4.20 & 4.20 & 4.20 & 4.20 \\
570: $\ell_2(\Delta\Psi^{600})/\ell_2(\Delta\Psi^{1200})$
571: & 5.00 & 5.00 & 5.00 & 5.00 \\
572: \hline \hline
573: \end{tabular}
574: \end{center}
575: \end{table}
576: %
577: grid resolution $K$ the truncation error is given by
578: %
579: \begin{align}
580: \Psi^K &= \Psi + \mathscr{O}\left(\frac{1}{K^2}\right),
581: \end{align}
582: %
583: where $\Psi$ is the exact and $\Psi^K$ the numerical solution. For simplicity
584: we will assume that the truncation error is either $-1/K^2$
585: or $+1/K^2$. If we use a reference solution obtained for $4K$ grid points
586: and compare solutions $\Psi^K$ and $\Psi^{2K}$ the ratio of the corresponding
587: $\ell_2$-norms becomes
588: %
589: \begin{align}
590: \left( \frac{\sum(\Psi_i^K-\Psi_i^{4K})^2}
591: {\sum(\Psi_i^{2K} - \Psi_i^{4K})^2}\right)^{1/2}
592: &= \left( \frac{\sum(\pm\frac{1}{K^2} \pm \frac{1}{16K^2})^2}
593: {\sum(\pm\frac{1}{4K^2} \pm \frac{1}{16K^2})^2} \right)^{1/2}
594: = \left| \frac{\pm16 \pm 1}{\pm4\pm1} \right| .
595: \end{align}
596: %
597: Considering the extreme cases, we expect a convergence factor between
598: 3 and $5\frac{2}{3}$. The truncation error of the high resolution
599: result will have significantly less influence on the comparison of
600: lower resolution results and the
601: factors should be closer to 4. Table \ref{convMink} shows our results for
602: the cosmic string in Minkowski space-time and clearly indicates
603: second order convergence.
604: In Fig.\,\ref{csmink} we show the string variables $P$ and $X$
605: for various values of $\alpha$ as a function of $w$.
606: %
607: \begin{figure}[t]
608: \centering
609: \epsfig{file=csmink.ps, height=350pt, width=200pt, angle=-90}
610: \caption{The cosmic string variables $P$ and $X$ are plotted for
611: $\alpha =$ 10, 1, 0.1, 0.01 (from ``left to right''). The two
612: families are labelled in the plot. As $\alpha$ increases,
613: both, $P$ and $X$ become more concentrated towards the origin.
614: Note that $w=3$ corresponds to $r\to \infty$ [cf. Eq.\,(\ref{w})].}
615: \label{csmink}
616: \end{figure}
617: %
618: Due to the rescaling (\ref{rescale_t})-(\ref{rescale_X}) the equations
619: for the cosmic string in Minkowski spacetime (\ref{mink_Prr}) and
620: (\ref{mink_Xrr}) do not explicitly contain the parameter $\eta$,
621: so the shape of the cosmic string fields expressed in terms of the rescaled
622: variables is independent of $\eta$.
623: Below we will see that this is no longer true
624: in curved spacetime where $\eta$, representing the mass of the string,
625: determines the strength of its coupling to gravity.
626: Fig.\,\ref{csmink} does, however, reveal a significant variation of the
627: profiles of the scalar and vector field
628: with the coupling ratio $\alpha$. As the scalar-vector
629: coupling becomes more dominant with respect to the self coupling of the
630: scalar field (larger $\alpha$), both $P$ and $X$ become more concentrated
631: towards the origin.
632:
633:
634: %========================================================================
635: \subsubsection{The static cosmic string coupled to gravity}
636: \label{CSstat}
637: %
638: %
639: The equations governing a static cosmic string in curved spacetime
640: are obtained from the general equations (\ref{nuur})-(\ref{xur}) by
641: setting all time derivatives to zero. If we combine first and second
642: spatial derivatives in a single operator as in equations (\ref{mink_Prr}),
643: (\ref{mink_Xrr}), we can write these equations as
644: %
645: \begin{align}
646: (r\nu_{,r})_{,r} =&\,\, -r\nu_{,r} \mu_{,r} + r\frac{\nu_{,r}^2
647: - \tau_{,r}^2}{\nu} + 8\pi\eta^2\left[ e^{-2\mu}
648: \nu^2 \frac{P_{,r}^2}{\alpha r}
649: - 2e^{2(\gamma+\mu)}r(X^2-1)^2\right], \label{nurr}
650: \\[10pt]
651: (r\tau_{,r})_{,r} =&\,\, 2r\frac{\tau_{,r} \nu_{,r}}{\nu}-r\tau_{,r}
652: \mu_{,r}\ , \label{taueq}
653: \end{align}
654: \begin{align}
655: (r^2\mu_{,r})_{,r} =&\,\, -r^2\mu_{,r}^2 - 8\pi \eta^2 \left[ e^{2\gamma}
656: X^2 P^2 + 2e^{2(\gamma+\mu)} \nu^{-1}
657: r^2 (X^2-1)^2\right], \\[10pt]
658: \gamma_{,r} =&\,\, \frac{r}{2(r\mu_{,r}+1)} \left[ \mu_{,rr}-\mu_{,r}^2
659: + \frac{1}{2\nu^2}\left(\tau_{,r}^2+ \nu_{,r}^2\right)
660: + 8\pi \eta^2 \left( X_{,r}^2 + \frac{1}{\alpha}
661: e^{-2\mu} \nu \frac{P_{,r}^2}{r^2} \right) \right],
662: \label{stat_gammar} \\[10pt]
663: \left( \frac{1}{r}P_{,r}\right)_{,r}
664: =&\,\, \frac{P_{,r}\mu_{,r}}{r} -\frac{P_{,r} \nu_{,r}}{r\nu}
665: +\alpha e^{2(\gamma+\mu)} \nu^{-1} \frac{PX^2}{r}, \\[10pt]
666: (rX_{,r})_{,r} =&\,\, -rX_{,r} \mu_{,r}
667: +X\left[ e^{2\gamma} \frac{P^2}{r}
668: + 4e^{2(\gamma+\mu)} \nu^{-1}r (X^2-1)\right]. \label{xrr}
669: \end{align}
670: %
671: After completing the code, we realised that in the
672: case of vanishing rotation $\tau$ the field equations
673: (\ref{nurr})-(\ref{xrr}) imply a simple relation between $\nu$,
674: $\mu$ and $\gamma$. An appropriate linear combination of these
675: equations and their spatial derivatives can be written as
676: %
677: \begin{align}
678: (\gamma + \mu - \ln{\nu})_{,rr} + \left(\frac{1}{r} + \mu_{,r}\right)
679: (\gamma + \mu - \ln{\nu})_{,r}
680: &= 0,
681: \end{align}
682: %
683: which after some manipulation becomes
684: %
685: \begin{align}
686: (\gamma + \mu - \ln{\nu})_{,r} &= C \frac{e^{-\mu}}{r}.
687: \end{align}
688: %
689: Here $C$ is a constant that has to vanish in order to ensure
690: finite derivatives at the origin.
691: In the static case we
692: adjust the functions $a_1$ and $a_2$ in the boundary conditions
693: (\ref{bound_nutr}), (\ref{bound_mutr}) so that $\nu=1$ and $\mu=0$
694: at the origin and consequently
695: %
696: \begin{align}
697: \gamma + \mu - \ln{\nu} &= 0, \label{gmn}
698: \end{align}
699: %
700: for all values of $r$. Even though $\tau$ will be zero in
701: the analysis in this section,
702: we will numerically solve the original system of
703: equations (\ref{nurr})-(\ref{xrr}) and
704: use (\ref{gmn}) as a test for the code. \\
705: In order to numerically solve the equations of a cosmic string coupled to
706: gravity, we rewrite them again as
707: a first order system. The differential operators appearing on the right hand
708: side suggest that we introduce the auxiliary
709: quantities $N=r\nu_{,r}$, $T=r\tau_{,r}$, $M=r^2\mu_{,r}$, $Q=r^{-1}P_{,r}$
710: and $R=rX_{,r}$. The system can then be written in the form
711: %
712: \begin{align}
713: \nu_{,r} &= \frac{N}{r}, \label{nur} \\[10pt]
714: \tau_{,r} &= \frac{T}{r}, \\[10pt]
715: \mu_{,r} &= \frac{M}{r^2}, \\[10pt]
716: P_{,r} &= rQ, \\[10pt]
717: X_{,r} &= \frac{R}{r}, \label{xr} \\[10pt]
718: N_{,r} &= \frac{N^2-T^2}{r\nu} - \frac{NM}{r^2}
719: - 16\pi \eta^2 e^{2\gamma + 2\mu} r(X^2-1)^2
720: + 8\pi \frac{\eta^2}{\alpha} e^{-2\mu} \nu^2 rQ^2, \\[10pt]
721: T_{,r} &= 2\frac{TN}{\nu r} - \frac{TM}{r^2}, \\[10pt]
722: M_{,r} &= -\frac{M^2}{r^2} - 8\pi\eta^2 e^{2\gamma} X^2P^2
723: - 16\pi\eta^2e^{2\gamma+2\mu} \nu^{-1}r^2 (X^2-1)^2, \\[10pt]
724: 2(r+M)\gamma_{,r} &= M_{,r} - 2\frac{M}{r} - \frac{M^2}{r^2}
725: + \frac{T^2+N^2}{2\nu^2} + 8\pi \eta^2 R^2
726: + 8\pi \frac{\eta^2}{\alpha} e^{-2\mu}\nu r^2Q^2, \\[10pt]
727: Q_{,r} &= \frac{QM}{r^2} - \frac{QN}{r\nu} +\alpha
728: e^{2\gamma+2\mu} \nu^{-1} \frac{PX^2}{r}, \\[10pt]
729: R_{,r} &= -\frac{RM}{r^2} + 4e^{2\gamma+2\mu} \nu^{-1} r X(X^2-1)
730: + e^{2\gamma} \frac{XP^2}{r}.
731: \end{align}
732: %
733: The corresponding equations in terms of the compactified radial coordinate $y$
734: are
735: %
736: \begin{align}
737: \nu_{,y} &= \frac{N}{y}, \\[10pt]
738: \tau_{,y} &= \frac{T}{y}, \label{CSTAT_TAUY} \\[10pt]
739: \mu_{,y} &= yM, \\[10pt]
740: P_{,y} &= \frac{Q}{y^5}, \\[10pt]
741: X_{,y} &= \frac{R}{y}, \\[10pt]
742: N_{,y} &= \frac{N^2-T^2}{y\nu} - yNM - 64\pi \eta^2 e^{2\gamma + 2\mu}
743: \frac{(X^2-1)^2}{y^5} + 8\pi \frac{\eta^2}{\alpha} e^{-2\mu} \nu^2
744: \frac{Q^2}{y^5}, \\[10pt]
745: T_{,y} &= 2\frac{TN}{y\nu} - yTM, \\[10pt]
746: M_{,y} &= -yM^2 - 32\pi \eta^2 e^{2\gamma} \frac{X^2 P^2}{y^3} - 64\pi \eta^2
747: e^{2\gamma + 2\mu} \nu^{-1} \frac{(X^2-1)^2}{y^7}, \\[10pt]
748: 2(y^2M-2) \gamma_{,y} = & \,\,\, y^2M_{,y}
749: + 4yM - y^3M^2 + \frac{N^2+T^2}{2y\nu^2}
750: +8\pi\eta^2 \frac{R^2}{y} + 8\pi \frac{\eta^2}{\alpha} e^{-2\mu}
751: \nu \frac{Q^2}{y^5},
752: \end{align}
753: \begin{align}
754: Q_{,y} &= yQM - \frac{QN}{y\nu} + 4\alpha e^{2\gamma+2\mu} \nu^{-1}
755: \frac{PX^2}{y}, \\[10pt]
756: R_{,y} &= -yRM + 4e^{2\gamma} \frac{XP^2}{y} + 16 e^{2\gamma+2\mu}\nu^{-1}
757: \frac{X(X^2-1)}{y^5}.
758: \end{align}
759: %
760: From the numerical point of view, the problem of solving these equations
761: is virtually identical to that of a static string in Minkowski spacetime.
762: The only difference is the much higher degree of complexity
763: of the equations due to the appearance of
764: $\nu$, $\tau$, $\mu$ and $\gamma$
765: as extra variables. We will discuss the numerical implementation of the
766: boundary conditions at the origin and at infinity in the next
767: section when we consider the case of a dynamic cosmic string. The
768: boundary conditions are given by
769: equations (\ref{bound_dynin}), (\ref{bound_dynout}). In the static case
770: we replace the conditions for $N$, $T$ and $M$ in (\ref{bound_dynin}) by
771: %
772: \begin{align}
773: \begin{split}
774: \nu = 1, \\[5pt]
775: \tau = 0, \\[5pt]
776: \mu = 0,
777: \end{split}
778: \label{stat_bound}
779: \end{align}
780: %
781: but otherwise use the same boundary conditions.
782: The solution is then obtained using the
783: relaxation method described in the previous section.
784: As our initial guess for the metric
785: variables we use Minkowskian values, and for the string variables $X$
786: and $P$ we use the previously calculated values for a Minkowskian string with
787: the same string parameters. Due to the appearance of $\tau$ or its derivatives
788: in all terms of (\ref{taueq}) the Geroch potential will
789: %
790: \begin{table}[t]
791: \caption{Convergence test for the static cosmic string in curved
792: space-time for $\alpha = 1$. The norm of the deviation
793: $\ell_2[\Delta \Psi^K]$ is defined by Eq.\,(\ref{CAUCHYNORM}).
794: As the grid resolution is increased, the deviation from the
795: high resolution result decreases quadratically to a good
796: approximation (see text for details).}
797: \begin{center}
798: \label{convStat}
799: \begin{tabular}{l|ccccc}
800: \hline \hline
801: & $\nu$ & $\mu$ & $\gamma$ & P $ X $ \\
802: \hline\\[-7pt]
803: $\ell_2(\Delta \Psi^{1200})$ & $1.28\cdot 10^{-7}$ & $2.51\cdot 10^{-6}$
804: & $2.39\cdot 10^{-6}$ & $5.95\cdot 10^{-7}$
805: & $4.16\cdot 10^{-7}$ \\
806: \hline\\[-7pt]
807: $\ell_2(\Delta\Psi^{150})/\ell_2(\Delta\Psi^{300})$
808: & 3.56 & 3.59 & 3.58 & 4.04 & 3.37 \\
809: $\ell_2(\Delta\Psi^{300})/\ell_2(\Delta\Psi^{600})$
810: & 3.76 & 3.79 & 3.78 & 4.19 & 3.60 \\
811: $\ell_2(\Delta\Psi^{600})/\ell_2(\Delta\Psi^{1200})$
812: & 4.58 & 4.61 & 4.60 & 4.98 & 4.44 \\
813: \hline \hline
814: \end{tabular}
815: \end{center}
816: \end{table}
817: %
818: stay zero in the relaxation
819: process and our solution has no rotation.\\
820: We have checked the code for convergence in the way described in section
821: \ref{SECcsmink}. We have chosen the unphysically large value
822: $\eta = 0.2$ here in order to guarantee convergence even for strong
823: coupling between matter and geometry. $\alpha$ is set to 1
824: as in the Minkowski case. The results are given in Table \ref{convStat}.
825: For convenience we only display the results for the fundamental variables
826: $\nu$, $\mu$, $\gamma$, $P$ and $X$. Since we do not incorporate rotation,
827: the result for $\tau$ is, as expected, exactly 0 and we do
828: not include it in Table \ref{convStat}.
829: Again the code is shown to be second order convergent.
830: %
831: \begin{figure}[b]
832: \centering
833: \epsfig{file=stat_ps.ps, height=400pt, width=283pt, angle=-90}
834: \caption{a) In the upper two panels we plot the string variables for
835: $\eta =$ 0.001 (dotted) and 0.2 (solid) as a function of
836: the radial variable $w$.
837: b) In the lower panels we have plotted the deviation from the
838: Minkowskian values rescaled by $\eta^2$ for $\eta = $
839: 0.001 (dotted), 0.01 (dashed), 0.1 (long dashed) and 0.2
840: (solid). Note that the curves for 0.001 and 0.01 almost exactly
841: coincide, which indicates the validity of the linear regime.
842: For larger $\eta$, however, the deviation shows a more
843: complicated behaviour.}
844: \label{stat_string}
845: \end{figure}
846: %
847: %
848: \begin{figure}[t]
849: \centering
850: \epsfig{file=stat_metr.ps, height=400pt, width=283pt, angle=-90}
851: \caption{The deviation of the metric variables $\nu$, $\mu$ and
852: $\gamma$ from Minkowskian values
853: rescaled by $\eta^2$ is plotted as a function of $w$
854: for $\eta=$ 0.001 (dotted), 0.01 (dashed), 0.1 (long dashed) and
855: 0.2 (solid). The dotted and the dashed curves almost exactly
856: coincide indicating the linear regime. As in the case of the
857: string variables we find a more complicated dependence for
858: $\eta \ge 0.1$. In the lower right panel we plot $\gamma$
859: $\mu$, $\ln{\nu}$ and their sum for $\eta=0.1$
860: which vanishes in accordance with Eq.\,(\ref{gmn})
861: to high accuracy.}
862: \label{stat_metric}
863: \end{figure}
864: %
865: In Fig.\,\ref{stat_string} and \ref{stat_metric} we plot the results obtained
866: for $N=2400$ grid points. In all these plots the relative coupling strength
867: is $\alpha = 1$, but qualitatively similar results are obtained for different
868: values of $\alpha$. We have already mentioned that the
869: % order of magnitude of the
870: effect of the string on the spacetime geometry is determined by
871: $\eta$. Therefore
872: we have compared the deviation of both the string variables and the metric
873: from the Minkowskian case for $\eta = $ 0.001, 0.01, 0.1 and 0.2.
874: In Fig.\,\ref{stat_string} we plot the string variables $P$ and $X$ for the
875: two extreme values and the deviation from the Minkowskian string
876: rescaled by $\eta^2$ for all four values. For small $\eta$ we
877: see that $\Delta P/\eta^2$ and $\Delta X/\eta^2$ is essentially independent of
878: $\eta$. In this case the deviation from Minkowskian values
879: can be treated as a small perturbation and
880: a linear dependence of $\Delta P$ and $\Delta X$ on $\eta^2$ is to be expected.
881: In the range $\eta =0.1 \ldots 0.2$ on the other hand,
882: we clearly leave the linear regime
883: and the deviation depends on $\eta$ in a much more complicated way.
884: These values, however, are 2 orders of magnitude larger than the value
885: $10^{-3}$ predicted
886: in current GUT theories (\citeNP{Vilenkin1994}).
887: The deviation of the
888: metric variables $\nu$, $\mu$ and $\gamma$ is plotted in the first three
889: panels of Fig.\,\ref{stat_metric}. Again we see the
890: linear behaviour for small $\eta$ and the transition to the non-linear
891: regime at $\eta \approx 0.1$. In the fourth panel of Fig.\,\ref{stat_metric}
892: we check Eq.\,(\ref{gmn}) for $\eta = 0.1$. We clearly see that
893: $\gamma + \mu + \ln{\nu}$ is approximately zero. Indeed (\ref{gmn}) is
894: satisfied to within $\approx 10^{-8}$ as compared with the order of
895: magnitude of the individual terms $10^{-1}$. \\
896:
897:
898: %=======================================================================
899: \subsubsection{The dynamic cosmic string}
900: \label{dynstring}
901: %
902: %
903: In the dynamic case all variables $\nu$, $\tau$, $\mu$, $\gamma$, $P$ and $X$
904: are functions of $u,r$ and we have to solve the system
905: (\ref{nuur})-(\ref{xur}) of partial differential equations.
906: In order to
907: control the behaviour of the solution at infinity, we need a
908: generalisation for PDEs of the relaxation scheme applied to ordinary
909: differential equations. In view of the characteristic feature of the
910: relaxation scheme, namely the simultaneous calculation of new function values
911: at all grid points, this generalisation leads directly to implicit
912: evolution schemes as used for hyperbolic or parabolic PDEs.
913: Therefore, the dynamic code is based on the implicit,
914: second order in space and time Crank-Nicholson scheme described in
915: section \ref{FDE_CN}. For this purpose we rewrite the
916: dynamic equations (\ref{nuur})-(\ref{xur}) as a first order system.
917: These equations involve radial
918: derivatives which may be written in terms of second order operators
919: exactly as in the static case (\ref{nurr})-(\ref{xrr}). This
920: naturally leads to the auxiliary quantities introduced in
921: Eqs.\,(\ref{nur})-(\ref{xr}). In terms of these variables the equations for
922: the dynamic cosmic string become
923: %
924: \begin{align}
925: \nu_{,r} =& \frac{N}{r}, \label{dyn_nur} \\[10pt]
926: \tau_{,r} =& \frac{T}{r}, \\[10pt]
927: \mu_{,r} =& \frac{M}{r^2}, \\[10pt]
928: P_{,r} =& rQ, \\[10pt]
929: X_{,r} =& \frac{R}{r}, \\[10pt]
930: \begin{split}
931: 2N_{,u} =& N_{,r} + \frac{T^2-N^2}{r\nu} + \frac{NM}{r^2}
932: + 2 \frac{\nu_{,u} N - \tau_{,u} T}{\nu} -\nu_{,u}
933: - \frac{\nu_{,u} M}{r} - N\mu_{,u} \\[10pt]
934: & + 8\pi\eta^2 \left[ 2e^{2(\gamma + \mu)} r(X^2-1)^2
935: + \frac{1}{\alpha} e^{-2\mu} \nu^2(2 P_{,u}Q - rQ^2)
936: \right],
937: \end{split} \\[10pt]
938: 2T_{,u} =& T_{,r} - 2\frac{TN}{r \nu} + 2\frac{\tau_{,u} N+\nu_{,u}T}{\nu}
939: + \frac{TM}{r^2} - \tau_{,u} - \frac{\tau_{,u} M}{r}
940: -T\mu_{,u}, \\[10pt]
941: 2M_{,u} =& M_{,r} + \frac{M^2}{r^2} - 2\mu_{,u} M - 2 r \mu_{,u}
942: + 8\pi\eta^2 \left[ e^{2\gamma} X^2P^2 +2
943: \frac{e^{2(\gamma+\mu)}}{\nu} r^2 (X^2-1)^2 \right], \\[10pt]
944: 2(r+M)\gamma_{,r} =& M_{,r} - 2\frac{M}{r} - \frac{M^2}{r^2}
945: + \frac{T^2+N^2}{2\nu^2} + 8\pi\eta^2 \left[ R^2
946: + \frac{1}{\alpha} e^{-2\mu}\nu r^2Q^2 \right],
947: \label{1gammar} \\[10pt]
948: 2Q_{,u} =& Q_{,r} - \frac{QM}{r^2} + Q \mu_{,u} - \frac{Q\nu_{,u}}{\nu}
949: - \frac{P_{,u} N}{r^2\nu} + \frac{QN}{r\nu} + \frac{P_{,u}}{r^2}
950: + \frac{P_{,u} M}{r^3} -\alpha \frac{e^{2(\gamma+\mu)}}{\nu}
951: \frac{PX^2}{r}, \\[10pt]
952: 2R_{,u} =& R_{,r} - X_{,u} - \frac{X_{,u} M}{r} + \frac{RM}{r^2}
953: - R\mu_{,u} - 4 \frac{e^{2(\gamma+\mu)}}{\nu} r X(X^2-1)
954: - e^{2\gamma} \frac{XP^2}{r} \label{Rur}.
955: \end{align}
956: %
957: The corresponding first order system in the outer region is given by
958: \begin{align}
959: \nu_{,y} =& -2 \frac{N}{y}, \label{nuy} \\[10pt]
960: \tau_{,y} =& -2 \frac{T}{y}, \\[10pt]
961: \mu_{,y} =& -2 yM, \\[10pt]
962: P_{,y} =& -2 \frac{Q}{y^5}, \\[10pt]
963: X_{,y} =& -2 \frac{R}{y}, \\[10pt]
964: \begin{split}
965: 2N_{,u} =& -\frac{1}{2}y^3 \left( N_{,y} -2yNM -2\frac{T^2-N^2}{y\nu}
966: \right) -y^2 \nu_{,u} M
967: +2\frac{\nu_{,u} N - \tau_{,u} T}{\nu}
968: \\[10pt]
969: & - N\mu_{,u} - \nu_{,u} +8\pi\eta^2 \left[2e^{2(\gamma + \mu)}
970: \frac{(X^2-1)^2}{y^2} +\frac{1}{\alpha} e^{-2\mu} \nu^2
971: \left( 2P_{,u} Q - \frac{Q^2}{y^2} \right) \right],
972: \end{split} \\[10pt]
973: 2T_{,u} =& -\frac{1}{2}y^3 \hspace{-0.025cm}
974: \left( T_{,y} -2y TM+4\frac{TN}{y\nu}\right) \hspace{-0.025cm}
975: - T\mu_{,u} -y^2 \tau_{,u} M
976: + 2\frac{\tau_{,u} N + \nu_{,u} T}{\nu} - \tau_{,u}, \\[10pt]
977: \begin{split}
978: 2M_{,u} =& -\frac{1}{2} y^3 (M_{,y} -2yM^2) - 2\frac{\mu_{,u}}{y^2}
979: -2\mu_{,u} M \\[10pt]
980: & + 8\pi\eta^2 \left[ e^{2\gamma} X^2P^2
981: + 2 \frac{e^{2(\gamma+\mu)}}{\nu} \frac{(X^2-1)^2}{y^4} \right],
982: \end{split} \\[10pt]
983: 2(y^2M+1) \gamma_{,y} =& \,\,\, y^2M_{,y} + 4yM +2y^3M^2
984: - \frac{N^2+T^2}{y\nu^2}
985: -16\pi\eta^2 \left[\frac{R^2}{y} + \frac{1}{\alpha} e^{-2\mu}
986: \nu \frac{Q^2}{y^5} \right], \\[10pt]
987: \begin{split}
988: 2Q_{,u} =& -\frac{1}{2}y^3 \left( Q_{,y} +2yQM -2\frac{QN}{y\nu} \right)
989: +y^4P_{,u} - y^4\frac{P_{,u} N}{\nu} - \frac{\nu_{,u} Q}{\nu}
990: \\[10pt]
991: & +y^6 P_{,u} M + Q\mu_{,u}
992: -\alpha e^{2(\gamma + \mu)} \nu^{-1} y^2PX^2,
993: \end{split} \\[10pt]
994: \begin{split}
995: 2R_{,u} =& -\frac{1}{2}y^3 (R_{,y} -2yRM) -X_{,u} -R\mu_{,u}
996: - y^2X_{,u} M - e^{2\gamma} y^2 XP^2 \\[10pt]
997: & - 4e^{2(\gamma+\mu)}\nu^{-1} \frac{X(X^2-1)}{y^2}.
998: \label{Ruy}
999: \end{split}
1000: \end{align}
1001: %
1002: The derivation of these equations and a number of other calculations
1003: in this work have been carried out with the algebraic computing package
1004: GRTensorII (\citeNP{Musgrave1996}).
1005: In order to solve these equations we must supplement them by
1006: appropriate initial and boundary conditions. We have already mentioned the
1007: boundary conditions on the axis (\ref{bound_nutr})-(\ref{bound_xtr}).
1008: In general we find that the dynamic code performs better
1009: if one imposes boundary conditions on the radial derivatives
1010: rather than the variables themselves. For the variables $\nu$, $\mu$, $\tau$,
1011: $P$ and $X$ we therefore impose the required
1012: boundary conditions on the initial data according to (\ref{mink_bound}) and
1013: (\ref{stat_bound}). In the subsequent
1014: evolution we impose the weaker condition that the radial derivatives
1015: $N$, $T$ and $R$ are finite on the axis. This ensures that the evolution
1016: equations propagate the axial conditions given on the initial
1017: data. For the variable $\mu$ we impose the condition that $M$ is zero
1018: on the axis which is equivalent to the rather weak condition that
1019: $r^2\mu_r$ vanishes there. The inverse power of $r$ in the
1020: definition of $Q$ makes it unsuitable to specify the value of this
1021: quantity at $r=0$ so in this case we work with the variable directly
1022: and require that $P=1$ on the axis.
1023: Finally the variable $\gamma$ is given by a purely radial equation and in
1024: this case we specify the value on the axis where $\gamma$ vanishes
1025: by virtue of Eq.\,(\ref{gammatr}). Therefore at $r=0$ we require
1026: \begin{align}
1027: \begin{split}
1028: N &= 0, \\
1029: T &= 0, \\
1030: M &= 0, \\
1031: \gamma &= 0, \\
1032: P &= 1, \\
1033: R &= 0.
1034: \end{split}
1035: \label{bound_dynin}
1036: \end{align}
1037: %
1038: For the boundary conditions at null infinity we know that regular
1039: solutions of the cylindrical wave equation have radial derivatives that decay
1040: faster than $1/r$ so that
1041: we may take the variables $N$, $T$, and $R$, which satisfy a wave
1042: type equation, to vanish at $y=0$. The asymptotics of $\mu$ are
1043: slightly different due to the additional power of $r$ in the radial
1044: derivative (similar to the spherically symmetric wave equation) but for a
1045: regular solution $\mu_{,y}$ vanishes at null infinity.
1046: The equation for $P$ does not satisfy a wave type equation due to the
1047: inverse power of $r$ but has
1048: asymptotic behaviour given by a modified Bessel function. The
1049: physically relevant finite solution has exponential decay so in this
1050: case one may impose the condition that $Q=0$ at $y=0$. Hence we
1051: require the solution to satisfy the following boundary conditions at $y=0$
1052: %
1053: \begin{align}
1054: \begin{split}
1055: N &= 0, \\
1056: T &= 0, \\
1057: \mu_{,y} &= 0, \\
1058: Q &= 0, \\
1059: R &= 0.
1060: \end{split}
1061: \label{bound_dynout}
1062: \end{align}
1063: %
1064: These boundary conditions are sufficient to determine the solution of
1065: the first order system (\ref{dyn_nur})-(\ref{Ruy}) while
1066: suppressing the unphysical
1067: solutions which are singular on the axis or null infinity. Note that
1068: $\gamma$ is determined by the constraint equation
1069: (\ref{gammar}), which is a first order ODE, and thus only needs one
1070: boundary condition. \\
1071: We finally note that all variables are related at the interface
1072: in the form $f_{K_1+1}=f_K$ analogous to Eqs.\,(\ref{int_P})-(\ref{int_R})
1073: in the case of a static string in Minkowski spacetime.
1074:
1075:
1076: %=======================================================================
1077: \subsection{Testing the dynamic code}
1078: \label{CStest}
1079: %
1080: %
1081: %%
1082: %\begin{figure}[t]
1083: % \centering
1084: % \epsfig{file=check_ww_delta.ps, height=400pt, width=150pt, angle=-90}
1085: % \caption{The deviation of the numerical $\nu$ and $\gamma$ from the
1086: % Weber-Wheeler solution as a function of $u$ and $w$ %(\ref{w})
1087: % obtained for 1920 grid points ($n_1=320$,
1088: % $n_2 = 1600$). The wave parameters are $a=2$, $b=0.5$.
1089: % Note that the error is amplified by $10^5$ and $10^7$ respectively.}
1090: %\label{plot_ww_delta}
1091: %\end{figure}
1092: %
1093: In this section we will describe four independent tests of the implicit code
1094: for the dynamic cosmic string, namely
1095: %
1096: \begin{list}{\rm{(\arabic{count})}}{\usecounter{count}
1097: \labelwidth1cm \leftmargin1.5cm \labelsep0.4cm \rightmargin1cm
1098: \parsep0.5ex plus0.2ex minus0.1ex \itemsep0ex plus0.2ex}
1099: \item reproducing the non-rotating vacuum solution of \lcite{Weber1957},
1100: \item reproducing the rotating vacuum solution of \scite{Xanthopoulos1986},
1101: \item using the results for the static cosmic string
1102: as initial data and verifying that the system stays in its static
1103: configuration,
1104: \item convergence analysis for the string hit by a Weber-Wheeler wave.
1105: \end{list}
1106: %
1107: Two additional tests arise in a natural way from the field equations and
1108: the numerical scheme. As described above there are two additional
1109: field equations which are consequences of the other equations.
1110: We have verified that these equations are satisfied to second order accuracy
1111: ($\sim \Delta r^2$). Furthermore the numerical scheme calculates the residuals
1112: of the algebraic equations to be solved, which have thus been monitored
1113: in test runs. They are satisfied to a much higher accuracy
1114: (double precision
1115: machine accuracy), so the total error is dominated by the truncation error
1116: of the second order differencing scheme. Another independent test is the
1117: comparison with the explicit CCM code which yields good
1118: agreement as long as the latter remains stable.
1119: The four main tests are now described in more detail.
1120:
1121:
1122: %========================================================================
1123: \subsubsection{The Weber-Wheeler wave}
1124: %
1125: %
1126: %
1127: \begin{figure}[t]
1128: \centering
1129: \epsfig{file=check_ww_delta.ps, height=400pt, width=200pt, angle=-90}
1130: \caption{The deviation of the numerical $\nu$ and $\gamma$ from the
1131: Weber-Wheeler solution as a function of $u$ and $w$ %(\ref{w})
1132: obtained for 1920 grid points ($K_1=320$,
1133: $K_2 = 1600$). The wave parameters are $a=2$, $b=0.5$.
1134: Note that the error is amplified by $10^5$ and $10^7$ respectively.}
1135: \label{plot_ww_delta}
1136: \end{figure}
1137: %
1138: %%
1139: %\begin{figure}[t]
1140: % \centering
1141: % \epsfig{file=check_as_delta1.ps, height=400pt, width=150pt, angle=-90}
1142: % \epsfig{file=check_as_delta2.ps, height=200pt, width=150pt, angle=-90}
1143: % \caption{The deviation of the numerical $\nu$, $\tau$, $\gamma$ from
1144: % Xanthopoulos' analytic solution as a function of $u$ and $w$
1145: % obtained for 1920 grid points
1146: % ($n_1=320$, $n_2 = 1600$).
1147: % Note that the error is amplified by $10^5$ and $10^6$
1148: % respectively.}
1149: %\label{plot_as_delta}
1150: %\end{figure}
1151: %%
1152: %%
1153: %\begin{figure}[t]
1154: % \centering
1155: % \epsfig{file=es_evol_nm.ps, height=400pt, width=150pt, angle=-90}
1156: % \epsfig{file=es_evol_gp.ps, height=400pt, width=150pt, angle=-90}
1157: % \epsfig{file=es_evol_s.ps, height=200pt, width=150pt, angle=-90}
1158: % \caption{The deviation of the metric and matter variables from the initial
1159: % data in the case of evolving a static cosmic string with
1160: % $\alpha=1$, $\eta=0.2$. For our standard grid with
1161: % $1920$ points, the
1162: % configuration stays static to an accuracy of $\approx 10^{-5}$
1163: % over a range of more than 30000 time steps.}
1164: % \label{evolstat}
1165: %\end{figure}
1166: %%
1167: In the first test we evolve the analytic solution given by \lcite{Weber1957},
1168: which describes a gravitational pulse of the ``+'' polarisation mode. We have
1169: given the analytic expressions in section \ref{CCM_ww} in terms of $t$, $r$
1170: [Eqs.\,(\ref{ww_X})-(\ref{ww_gamma})] and in terms of $u$, $y$
1171: [Eqs.\,(\ref{ww_yX})-(\ref{ww_ygamma})].
1172: The corresponding equations in characteristic
1173: coordinates $u$, $r$ in the inner region
1174: are easily obtained from the coordinate transformation $t=u+r$.
1175: We prescribe $\nu$ as initial data according
1176: to the analytic expressions obtained for $a=2$ and $b=0.5$ and set the other
1177: free variables to zero, while $\gamma$ is calculated
1178: via quadrature from the constraint equation (\ref{gammar}).
1179: In Fig.\,\ref{plot_ww_delta} we show
1180: the deviation of the numerical results from the analytic one for
1181: $K=1920$ grid points (320 points in the inner, 1600
1182: points in the outer region) and a Courant factor of 0.5 with respect to
1183: the inner region. The convergence analysis (see below)
1184: shows that this number of points provides sufficient resolution in the outer
1185: region while still keeping computation times at a tolerable level.
1186: All computations presented in the remainder of section \ref{cstr} have
1187: been obtained with these grid parameters, unless stated otherwise.
1188: The code stays stable for much longer time intervals than shown in the figure,
1189: but does not reveal any further interesting features as the analytic solution
1190: approaches its Minkowskian values and the error goes to zero.
1191:
1192:
1193: %======================================
1194: \subsubsection{The rotating solution of Xanthopoulos}
1195: %
1196: %
1197: %
1198: \begin{figure}[t]
1199: \centering
1200: \epsfig{file=check_as_delta1.ps, height=400pt, width=175pt, angle=-90}
1201: \epsfig{file=check_as_delta2.ps, height=200pt, width=175pt, angle=-90}
1202: \caption{The deviation of the numerical $\nu$, $\tau$, $\gamma$ from
1203: Xanthopoulos' analytic solution as a function of $u$ and $w$
1204: obtained for 1920 grid points
1205: ($K_1=320$, $K_2 = 1600$).
1206: Note that the error is amplified by $10^5$ and $10^6$
1207: respectively.}
1208: \label{plot_as_delta}
1209: \end{figure}
1210: %
1211: \scite{Xanthopoulos1986} has derived an analytic vacuum solution for Einstein's
1212: field equations in cylindrical symmetry containing both the ``+'' and
1213: ``$\times$'' polarisation mode. Its analytic form in terms of
1214: our metric variables is given by Eqs.\,(\ref{as_Q})-(\ref{as_ygamma})
1215: in section \ref{as}. Again
1216: the transformation to coordinates $u$, $r$ in the inner region
1217: results straightforwardly from $t=u+r$. The solution has one free parameter
1218: $a$ set to one in this calculation.
1219: The error of our numerical results is displayed in Fig.\,\ref{plot_as_delta},
1220: where we have used the same grid parameters as in the Weber-Wheeler case.
1221: Again we have run
1222: the code for longer times and found that the error approaches zero.
1223: We conclude that the code reproduces both analytic vacuum solutions
1224: with excellent accuracy comparable to that of the CCM code
1225: and exhibits long term stability.
1226:
1227:
1228: %========================================
1229: \subsubsection{Evolution of the static cosmic string}
1230: %
1231: %
1232: %
1233: \begin{figure}[t]
1234: \centering
1235: \epsfig{file=es_evol_nm.ps, height=400pt, width=175pt, angle=-90}
1236: \epsfig{file=es_evol_gp.ps, height=400pt, width=175pt, angle=-90}
1237: \epsfig{file=es_evol_s.ps, height=200pt, width=175pt, angle=-90}
1238: \caption{The deviation of the metric and matter variables from the initial
1239: data in the case of evolving a static cosmic string with
1240: $\alpha=1$, $\eta=0.2$. For our standard grid with
1241: $1920$ points, the
1242: configuration stays static to an accuracy of $\approx 10^{-5}$
1243: over a range of more than 30000 time steps.}
1244: \label{evolstat}
1245: \end{figure}
1246: %
1247: The tests described above only involve vacuum solutions, so the matter
1248: part of the
1249: code and the interaction between matter and geometry has not been tested.
1250: An obvious test involving matter and geometry is to use the result for the
1251: static cosmic string in curved spacetime as initial
1252: data and evolve this scenario. All variables should, of course, remain
1253: at their initial values. We have evolved the static string data for
1254: our standard grid and the parameter set, $\alpha = 1$ and $\eta=0.2$,
1255: which corresponds to a strong back-reaction of the string on the
1256: metric. The results are shown in Fig.\,\ref{evolstat}. The system stays
1257: in its static configuration with high accuracy over a time interval
1258: which is significantly longer than the dynamically relevant
1259: timescale of the vacuum solutions discussed above.
1260:
1261:
1262: %=======================================
1263: \subsubsection{Convergence analysis}
1264: %
1265: %
1266: %
1267: \begin{figure}[t]
1268: \centering
1269: \epsfig{file=conv_cs_nmgp.ps, height=400pt, width=283pt, angle=-90}
1270: \epsfig{file=conv_cs_s.ps, height=200pt, width=141pt, angle=-90}
1271: \caption{The convergence factor $\ell_2[\Psi^{1920}] / \ell_2[\Psi^{2880}]$
1272: is plotted as a function of $u$.
1273: We expect a convergence factor of 2.25 since the number of
1274: grid points is multiplied by 1.5.
1275: Even though our results show weak variability at later times,
1276: second order convergence is maintained
1277: throughout long runs (more than 30000 time steps with $K=1920$).}
1278: \label{cs_conv}
1279: \end{figure}
1280: %
1281: Our investigation of the interaction between the cosmic string and gravitational
1282: waves will focus on the string being hit by a wave of the Weber-Wheeler type.
1283: In order to check this scenario for convergence we have run
1284: the code for the parameter set $\eta=0.2$, $\alpha=1$, $a=2$, $b=0.5$
1285: for different grid resolutions, where $a$ and $b$
1286: are again the width and amplitude
1287: of the Weber-Wheeler wave. In our case it is of particular interest to
1288: investigate the time dependence of the convergence to see what resolution we
1289: need in order to obtain reliable results for long runs.
1290: We calculate the convergence rate again according to equation (\ref{l2norm}).
1291: The high
1292: resolution reference solution has been calculated for $K=4320$ grid points.
1293: In Fig.\,\ref{cs_conv} we show the convergence factor
1294: $\ell_2[\Psi^{1920}]/\ell_2[\Psi^{2880}]$ as a function of $u$
1295: for $\nu$, $\mu$, $\gamma$, $P$ and $X$. The initial data for $\tau$ is
1296: identically zero for this scenario and stays zero during the evolution.
1297: The number of grid points is increased by a factor of 1.5 here (instead of the
1298: more commonly used 2) to reduce the computation time.
1299: Only points common to all grids have been used in the
1300: sum in equation (\ref{l2norm}). For second order convergence we would expect
1301: a convergence factor of $1.5^2$. Although the results in
1302: Fig.\,\ref{cs_conv}
1303: show weak variations with $u$, second order convergence is clearly maintained
1304: for long runs.
1305: In each case the outer region contains 5 times as many grid points as
1306: the inner region (e.g. $K_1=320$, $K_2=1600$ for the $K=1920$ case). The
1307: reason for this is that in the dynamic evolutions
1308: $X$ and especially $P$ exhibit significant spatial variations
1309: out to large radii. Due to the compactification, the spatial resolution of
1310: our grid decreases towards null infinity and in order
1311: to resolve the spatial
1312: variations of the string variables out to sufficiently large radii we
1313: therefore have
1314: to introduce a large number of grid points in the outer region.
1315: No such problems occur in the inner region. If significantly
1316: fewer grid points are used in the outer region for this analysis, the
1317: convergence properties of the string variables can deteriorate
1318: to roughly first order level.
1319:
1320:
1321: %=========================================================================
1322: \subsection{Time dependence of the string variables}
1323: \label{CSevol}
1324: %
1325: %
1326: \subsubsection{Static cosmic strings excited by gravitational waves}
1327: %
1328: \begin{figure}[t]
1329: \centering
1330: \epsfig{file=ini_nps0001.ps, height=350pt, width=200pt, angle=-90}
1331: \caption{The initial data for $\nu$, $P$ and $X$ at $u_0=-20$
1332: for the standard parameters $\alpha=1$, $\eta=0.001$,
1333: $a=2$, $b=0.5$.
1334: The gravitational wave pulse is located in a region
1335: where the string fields $P$ and $X$ have almost fallen off
1336: to their asymptotic values.}
1337: \label{ini_nps0001}
1338: \end{figure}
1339: %
1340: %
1341: \begin{figure}[t]
1342: \centering
1343: \epsfig{file=mplot_nmgx.ps, height=450pt, width=300pt, angle=-90}
1344: \epsfig{file=mplot_p.ps, height=450pt, width=150pt, angle=-90}
1345: \caption{The metric and string variables are plotted as functions
1346: of $w$ at $u=-20$ (dotted), $u=-10$ (long dotted),
1347: $u=0$ (dashed), $u=2$ (long dashed) and $u=10$ (solid line).
1348: For clarity the graphs of $P$ are distributed over two panels.
1349: The wave pulse (in $\nu$) initially moves inwards. It excites
1350: the string, is reflected at the origin and travels outwards.
1351: After $u=10$ only $P$ differs significantly from the
1352: static configuration as the oscillations
1353: slowly decay and propagate towards larger
1354: radii (cf. Fig.\,\ref{ringing}).}
1355: \label{mplot}
1356: \end{figure}
1357: %
1358: %
1359: \begin{figure}[t]
1360: \centering
1361: \epsfig{file=ringing.ps, height=450pt, width=200pt, angle=-90}
1362: \caption{The cosmic string variable $P$ is shown as a function
1363: of radius and time for $\alpha = 0.2$ (left) and
1364: $\alpha = 1$ (right) (all other parameters have standard
1365: values). Note that we use the radial
1366: variable $r$ out to $r=50$ here. The ringing can clearly
1367: be seen and shows a lower frequency for smaller $\alpha$.}
1368: \label{ringing}
1369: \end{figure}
1370: %
1371: The scenario we are going to investigate in this section is an initially
1372: static cosmic string hit by a gravitational wave of
1373: Weber-Wheeler type. For this purpose we use the static results with
1374: two modifications as initial data.
1375: Firstly the static metric function $\nu_0$ is multiplied
1376: by the exact Weber-Wheeler solution to simulate the gravitational wave
1377: pulse. Thus we guarantee that initially the cosmic string is indeed
1378: in an equilibrium configuration provided the wave pulse is located
1379: sufficiently far away from the origin and its interaction with the string
1380: is negligible.
1381: %If the pulse is located sufficiently far away from the origin
1382: %its interaction with the cosmic string is negligible and the initial
1383: %string configuration is indeed static.
1384: Ideally the numerical calculation would start with the incoming wave located
1385: at past null infinity. In order to approximate this scenario, we
1386: found it was sufficient to use the large negative initial time $u_0=-20$.
1387: The second modification is to calculate $\gamma$ from the constraint
1388: equation (\ref{gammar}) to preserve consistency with the Einstein field
1389: equations.
1390: In Fig.\,\ref{ini_nps0001}
1391: the corresponding initial data for $\nu$, $P$ and $X$ are shown for
1392: the parameter set $\eta=0.001$, $\alpha=1$, $a=2$ and $b=0.5$.
1393: From now on we will refer to these values as ``standard parameters''
1394: and only specify parameters if they take on non-standard values.
1395: %Note that $\tau$ vanishes on the initial slice in this case and stays
1396: %identically zero throughout the evolution.
1397: % The case of rotating gravitational
1398: % waves hitting a cosmic string will be analysed in a future publication.
1399: The time evolution of the ``standard configuration'' is shown in
1400: Fig.\,\ref{mplot} where we plot $\nu$, $\mu$, $\gamma$, $P$ and $X$
1401: as functions of $w$ at times $-20$, $-10$, $0$, $2$ and $10$. While the wave
1402: pulse is still far away from the origin, its interaction with the cosmic
1403: string is negligible (dotted lines).
1404: When it reaches the core region it excites
1405: the cosmic string and the scalar and vector field start oscillating (dashed
1406: curves).
1407: After being reflected at the origin, the wave pulse travels along the
1408: outgoing characteristics and the metric variables
1409: $\nu$, $\mu$ and $\gamma$ quickly settle down into their static configuration
1410: which is close to Minkowskian values for $\eta = 10^{-3}$.
1411: The vector and scalar field of the cosmic string, on the other hand,
1412: continue ringing albeit with a different character. Whereas the
1413: oscillations of the scalar field $X$ are dominant in the range $r\le 2$
1414: and have significantly decayed at $u=10$ as shown in the figure,
1415: the vector field oscillations propagate to large radii and fall off very
1416: slowly (solid curves). This behaviour is also
1417: illustrated in the right panel of Fig.\,\ref{ringing} which shows
1418: a contour plot of $P$ as a function of $(u,r)$ out to $r=50$.
1419: We shall see below, that the oscillations of
1420: $P$ will also decay eventually and the
1421: cosmic string will asymptotically settle back into its equilibrium
1422: configuration.
1423:
1424:
1425: %=====================================
1426: \subsubsection{Frequency analysis}
1427: %
1428: %
1429: We will now quantitatively analyse the oscillations of the scalar and
1430: vector field of the cosmic string. Since we are working in rescaled
1431: coordinates, physical time and distance are
1432: measured in units of $1/\sqrt{\lambda} \eta$ and
1433: frequency in its inverse. To avoid complicated notation, however,
1434: we will omit the units from now on unless the meaning is unclear.
1435: In order to measure frequencies, we Fourier analyse the time evolution
1436: of the scalar and vector field for fixed radius $r$. Fig.\,\ref{fourier}
1437: shows $P$ and $X$ for standard parameters
1438: as functions of time at $r=1$ together with the corresponding
1439: power spectra. In each Fourier spectrum we can see three peaks. Those very
1440: close to $f=0$ are merely due to the offset of the data and the gradual
1441: change of the fields over long times. We therefore do not attribute these
1442: peaks to the oscillations of the fields and will not consider them in the
1443: ensuing discussion. We have calculated similar
1444: power spectra for a large class of parameter sets and in most cases
1445: found two peaks at non-zero frequencies. In order to interprete
1446: the frequencies, it is convenient to plot them as functions of the
1447: relative coupling strength $\alpha$. The result is shown in
1448: Fig.\,\ref{f_alpha}. In this figure the solid lines show the frequency values
1449: calculated for the scalar and vector field from the linearised equations
1450: (see \citeNP{Sjodin2001})
1451: %
1452: \begin{align}
1453: f_X &= \frac{\sqrt{2}}{\pi}, \label{CSTR_LINFX} \\
1454: f_P &= \frac{\sqrt{\alpha}}{2\pi}. \label{CSTR_LINFP}
1455: \end{align}
1456: %
1457: We can thus associate the
1458: constant frequency $f=0.45$ with the scalar field $X$ and the
1459: $\alpha$ dependent frequency with the vector field $P$.
1460: We will refer to these frequencies as $f_X$ and $f_P$ from now on.
1461: The $\alpha$-dependency of $f_P$ is also illustrated in Fig.\,\ref{ringing}
1462: where we show contour plots of $P$ obtained for $\alpha=0.2$ and $\alpha=1$.
1463: The oscillation frequency is significantly larger for $\alpha=1$. \\
1464: In Fig.\,\ref{f_alpha} we can see that the
1465: frequencies associated with the scalar
1466: and vector field become similar near $\alpha=8$. For this value
1467: it can be shown that the masses associated with the scalar and vector field
1468: become equal (see for example \citeNP{Sjodin2001}). The frequencies are
1469: difficult to resolve in these cases and we have only found one peak in the
1470: Fourier spectra. The resulting values are shown as filled lozenges in
1471: the figure. In this context it is worth mentioning that
1472: the accuracy of the measurements of the
1473: frequencies is limited by the resolution of the Fourier spectra
1474: which again is limited by the time interval covered in the evolution.
1475: In Fig.\,\ref{fourier} we can see however, that the oscillations of
1476: both $P$ and $X$ gradually die out in later stages of the evolutions,
1477: so that it becomes increasingly difficult to extract more information
1478: about the frequencies by using larger integration times.
1479: The evolutions used for this analysis provide an
1480: accuracy $\Delta f \approx 0.01$ which corresponds approximately to one bin
1481: in the frequency spectra. \\
1482: It is interesting
1483: to see that in the non-linear evolution the distinction between
1484: the oscillations of the vector and the scalar field is not as clear
1485: as in the linear case which is demonstrated by the presence of
1486: two peaks in the Fourier spectra.
1487: We attribute this feature to the interaction between the scalar and vector
1488: component of the cosmic string. Concerning the radial dependence
1489: of the spectra we have in general found that the
1490: characteristic mode of $X$ resulted in stronger peaks at smaller radius,
1491: that of $P$ was stronger at larger radii.
1492: This variation of the relative strength of
1493: the oscillations with radius
1494: confirms the corresponding observation in Fig.\,\ref{mplot}.
1495: %
1496: \begin{figure}[t]
1497: \centering
1498: \epsfig{file=fourier.ps, height=400pt, width=283pt, angle=-90}
1499: \caption{Upper panels: The variables $P$ and $X$ at $r=1$ are plotted
1500: as functions of $u$ for $\alpha=1$, $\eta=0.001$, $a=2$ and $b=0.5$.
1501: Lower panels: The corresponding power
1502: spectra show three peaks each. That near $f=0$ is merely due
1503: to a constant offset and the variation of the fields on
1504: long time scales and thus not associated with the oscillations.
1505: From the linear equations one can infer that the peaks
1506: at $f=0.45$ can be identified with the oscillation of the
1507: scalar field, the peaks at $f=0.16$ with those of the vector
1508: field. Note that due to our rescaling of the coordinates,
1509: $u$ is measured in units of $1/\sqrt{\lambda} \eta$.}
1510: \label{fourier}
1511: \end{figure}
1512: %
1513: In order to investigate the dependency of the oscillations on
1514: %
1515: \begin{figure}[t]
1516: \centering
1517: \epsfig{file=fpx_alpha.ps, height=300pt, width=200pt, angle=-90}
1518: \caption{The frequencies obtained from the Fourier analysis of the
1519: oscillations of the scalar and vector fields
1520: are plotted as functions of $\alpha$.
1521: The curves show the frequencies of the scalar and vector field
1522: predicted by an analysis of the linearised equations.
1523: For $5 \le \alpha \le 8$ the predicted values for
1524: $P$ and $X$ are similar and we find only one frequency.
1525: These values are plotted as filled lozenges.}
1526: \label{f_alpha}
1527: \end{figure}
1528: %
1529: $\eta$, $a$, $b$ and the radial position $r$, we have varied each
1530: parameter over at least two orders of magnitude while keeping the other
1531: parameters at standard values. We have found the following dependencies:
1532: %
1533: \begin{list}{\rm{(\arabic{count})}}{\usecounter{count}
1534: \labelwidth1cm \leftmargin1.5cm \labelsep0.4cm \rightmargin1cm
1535: \parsep0.5ex plus0.2ex minus0.1ex \itemsep0ex plus0.2ex}
1536: \item The frequencies of both $X$ and $P$ do not show any variations
1537: with $\eta$ for $\eta < 0.1$. (Note that $\eta$ does, however,
1538: appear in the units). For larger values of $\eta$, the non-linear
1539: interaction between string and geometry becomes dominant and we did
1540: not detect a simple relation between frequency maxima and parameters.
1541: \item The variation of the parameters $a$ and $b$,
1542: the width and amplitude of the Weber-Wheeler pulse, has
1543: no measurable effect on the frequencies of $X$ and $P$, but
1544: only determined the amplitude of the oscillations. A narrow, strong pulse
1545: leads to larger amplitudes.
1546: \item For small $r$ the oscillations in $X$ are stronger, whereas
1547: those for $P$ dominate at large $r$. The frequency values, however,
1548: do not depend on the radius. For radii greater than 10
1549: the oscillations
1550: in $X$ had decayed so strongly that we could no longer measure its
1551: frequency.
1552: \end{list}
1553: %
1554: We have also checked the empirical relation between the coupling constant
1555: $\alpha$ and the frequencies $f_X$ and $f_P$. For this purpose we
1556: have performed a linear regression analysis of the double logarithmic
1557: data of Fig.\,\ref{f_alpha} excluding the cases where only one
1558: frequency is observed. We obtain power law indices $\sigma_X = 0.00$
1559: and $\sigma_P = 0.50$, so that
1560: %
1561: \begin{align}
1562: f_X & \sim {\rm const}, \label{fX}\\
1563: f_P & \sim \sqrt{\alpha}, \label{fP}
1564: \end{align}
1565: %
1566: which agrees with Eqs.\,(\ref{CSTR_LINFX}), (\ref{CSTR_LINFP}).
1567: If we transform this result back
1568: into physical units using $\alpha=e^2/\lambda$, we arrive at the following
1569: relations for the physical variables
1570: %
1571: \begin{align}
1572: f_X & \sim \sqrt{\lambda} \eta, \\[10pt]
1573: f_P & \sim e \eta.
1574: \end{align}
1575: %
1576: As shown in \lcite{Shellard1994} up to constant factors
1577: $\sqrt{\lambda} \eta$ and $e \eta$ are the masses
1578: of the scalar and the vector field, $m_X$ and $m_P$, so that
1579: $X$ and $P$ have characteristic frequencies
1580: %
1581: \begin{align}
1582: f_X & \sim m_X, \\[10pt]
1583: f_P & \sim m_P.
1584: \end{align}
1585: %
1586: Since the frequencies for $X$ and $P$ seem only to depend upon the
1587: respective masses we have attempted to confirm these results by
1588: considering the oscillations of a cosmic string in two further scenarios.
1589: Firstly since the frequencies do not depend upon the shape of the Weber-Wheeler
1590: %
1591: \begin{figure}[b]
1592: \centering
1593: \epsfig{file=settle_nm.ps, height=400pt, width=175pt, angle=-90}
1594: \epsfig{file=settle_gs.ps, height=400pt, width=175pt, angle=-90}
1595: \epsfig{file=settle_p.ps, height=250pt, width=150pt, angle=-90}
1596:
1597: \vspace{0.3cm}
1598: \caption{The upper four plots show the difference between the evolved
1599: functions $\nu$, $\mu$, $\gamma$ and $X$ and their corresponding
1600: static results. For $P$ a similar 3-dimensional plot is not suitable
1601: since it fails to resolve the oscillations of the vector field.
1602: Therefore we plot the $\ell_2$-norm (dashed line)
1603: and the maximum (solid line) of $\Delta P$ as
1604: a function of time. $\nu$, $\mu$, $\gamma$ and $X$ quickly settle
1605: down in their equilibrium configuration to numerical accuracy.
1606: The decay of the oscillations of $P$ takes significantly more
1607: time but eventually $P$ also approaches its equilibrium state.}
1608: \label{settle}
1609: \end{figure}
1610: %
1611: pulse we take as initial data the static values for the metric
1612: variables but excite the string by adding a Gaussian perturbation to
1613: either the $X$ or $P$ static initial values. The evolution is then
1614: computed using the fully coupled system. Secondly since the
1615: frequencies do not seem to depend upon the strength of the coupling to
1616: the gravitational field we have completely decoupled the gravitational
1617: field and considered the evolution of a cosmic string in Minkowski
1618: spacetime. The initial data is taken to be that for a static string in
1619: Minkowski spacetime with a Gaussian perturbation to either the $X$ or
1620: $P$ values. The evolution is then computed using the equations for a
1621: dynamical string in a Minkowskian background [equations (\ref{pur}) and
1622: (\ref{xur}) with the metric variables set to Minkowskian values].
1623: In both cases we obtain the same frequencies, to within an
1624: amount $\Delta f = 0.01$, that we find in the original case of the
1625: fully coupled system excited by a Weber-Wheeler pulse. Furthermore the
1626: frequencies did not depend on the location or shape of the field perturbation
1627: nor upon the choice of $X$ or $P$ as the perturbed field.
1628:
1629:
1630: %========================================================================
1631: \subsubsection{The long term behaviour of the dynamic string}
1632: %
1633: %
1634: The time evolution shown in Figs.\,\ref{mplot} and \ref{ringing} indicates
1635: that the oscillations of the cosmic string excited by gravitational waves
1636: gradually decay and metric and string settle down into an equilibrium state.
1637: In order to investigate the long term behaviour in detail
1638: we have evolved the variables for a much longer time
1639: ($-20\le u \le 410$) than in the numerical evolutions discussed above.
1640: The unphysically large value of $\eta = 0.1$ is chosen for this
1641: calculation in order to guarantee a
1642: strong interaction between spacetime geometry and the cosmic string.
1643: In Fig.\,\ref{settle} we show the difference $\Delta f :=
1644: f_{\rm evol} - f_{\rm stat}$
1645: between the time-dependent $\nu$, $\mu$, $\gamma$ and $X$ and their
1646: corresponding static results obtained for the same parameters.
1647: For the vector field $P$ a similar 3-dimensional plot would require
1648: an extreme resolution to properly display the oscillations of the vector
1649: field (cf.\ Fig.\,\ref{mplot}). For this reason we proceed differently
1650: and calculate the $\ell_2$-norm and the maximum of $\Delta P$ for each
1651: slice $u={\rm const}$. Both functions
1652: are plotted versus time in Fig.\,\ref{settle}.
1653: The incoming wave pulse can clearly be seen as a strong deviation of $\nu$
1654: from the static function. The pulse excites the cosmic
1655: string and is reflected at the origin at $u=0$. The metric variables and
1656: the scalar field $X$ then quickly reach their equilibrium values. The
1657: oscillations in $P$ decay on a significantly longer time scale
1658: which is also evident in Figs.\,\ref{mplot} and \ref{ringing}
1659: and the $\ell_2$-norm of $\Delta P$ slowly approaches 0.
1660: Significantly longer runs than shown here are prohibited by the required
1661: computation time, but the results indicate that $P$ will also eventually
1662: reach its equilibrium configuration.
1663:
1664:
1665: %========================================================================
1666: \subsection{Discussion}
1667: %
1668: %
1669: In the previous two sections we have studied numerical problems
1670: in cylindrical symmetry with particular emphasis on the use of
1671: characteristic methods and the compactification of spacetime. \\
1672: This work has completed the 1-dimensional stage of the
1673: Southampton Cauchy-characteristic matching project by presenting
1674: for the first time a long-term stable second order convergent code
1675: for general cylindrically symmetric vacuum spacetimes with both the
1676: $+$ and $\times$ polarisation mode. In order to obtain long-term
1677: stability it was crucial to formulate the problem in a way that
1678: simplifies the relations at the interface where information
1679: is transferred between the interior Cauchy and the exterior
1680: characteristic region. In this particular case we achieved the
1681: simplification by applying the Geroch decomposition to both regions
1682: which contrasts with the less successful previous attempts where
1683: the Killing direction was factored out in the exterior region only.
1684: In view of the numerical subtleties involved with the interpolation
1685: techniques at the interface the importance of a suitable
1686: choice of variables may not be too surprising. Nevertheless we stress
1687: the significance of this result concerning Cauchy-characteristic
1688: matching codes in higher dimensions. The structure of the null geodesics
1689: will inevitably become much more complicated if the restriction of
1690: cylindrical symmetry is dropped and the physical variables
1691: are allowed to depend on the angular coordinates. Correspondingly the
1692: transformation between variables at the interface will also be more
1693: complicated. In view of our results it seems important to carefully
1694: choose the variables describing the spacetime in both regions and
1695: aim for ``simple'' transformation laws. \\
1696: The inclusion of matter in the form of cosmic strings resulted in
1697: qualitatively new numerical problems that finally were solved by the
1698: use of specially adapted numerical methods. The incorporation of
1699: null infinity proved to be important here for the specification
1700: of outer boundary conditions on the matter variables. It was the existence
1701: of unphysical exponentially diverging solutions of the equations
1702: for a cosmic string that required a special numerical treatment.
1703: We were able to suppress the unphysical diverging solutions by
1704: solving the equations for a cosmic string with
1705: a relaxation scheme in the static case and an
1706: implicit evolution scheme in the dynamic case. We have thus been able
1707: to develop the first fully non-linear simulation of a static and
1708: a dynamic cosmic string
1709: coupled to gravity which implements the exact boundary conditions
1710: at both the origin and infinity. The resulting codes have been used
1711: to study the interaction between a cosmic string and a gravitational wave
1712: pulse. We have found that the gravitational wave pulse excites the
1713: cosmic string which then starts oscillating with frequencies proportional
1714: to the particle masses associated with the scalar and vector field.
1715: The same frequencies have been observed if we excite the cosmic
1716: string with a Gaussian perturbation to the scalar or vector field. \\
1717: From a numerical point of view an interesting result of the numerical
1718: solution of the equations for a dynamic cosmic string
1719: concerns the transfer of information at the interface. We
1720: have illustrated this in Fig.\,\ref{grid} where two grid points
1721: $K_1$, $K_1+1$ have been used for the spatial position $r=1$. The grid
1722: point $K_1$ contains the variables of the interior region at $r=1$, whereas
1723: the variables of the exterior region are specified at the same position on grid
1724: point $K_1+1$. The corresponding implementation of the interface
1725: is remarkably simple as illustrated by
1726: Eqs.\,(\ref{int_P})-(\ref{int_R}) which represent the interface for the
1727: static cosmic string in Minkowski spacetime. The corresponding
1728: equations in the dynamic case coupled to gravity are equally trivial.
1729: Even if different variables are used in the interior and exterior
1730: region, one is still able to transform the variables locally at
1731: the grid points $K_1$ and $K_1+1$ and there is no need to use
1732: interpolation techniques as in the case of the explicit numerical methods
1733: used in section \ref{ccm}. We attribute the possibility of using this
1734: simple implementation of the interface to the fact that all
1735: function values are calculated simultaneously on the new time slice
1736: in an implicit scheme,
1737: so that there is no hierarchical order according to which the
1738: function values have to be calculated. We have probed such a ``local
1739: interface'' in an implicit Cauchy-characteristic matching code
1740: for cylindrically symmetric, non-rotating vacuum spacetimes and
1741: achieved a long term stable evolution with an accuracy comparable to the
1742: explicit code described in section \ref{ccm}. Even though an interface
1743: based on interpolation performs satisfactorily in cylindrical symmetry
1744: this may no longer be the case in higher dimensional problems where
1745: the interpolation techniques will be significantly more complicated.
1746: On the other hand we can see no obvious reason why a
1747: ``local interface'' in
1748: combination with an implicit numerical scheme should differ significantly
1749: from that used in the 1-dimensional case.
1750:
1751:
1752:
1753:
1754:
1755:
1756: