gr-qc0208009/css.tex
1: \documentclass[12pt]{article}
2: \addtolength{\textheight}{2cm}
3: \addtolength{\hoffset}{-0.75cm}
4: \addtolength{\voffset}{-1cm}
5: \addtolength{\textwidth}{1.5cm}
6: \usepackage{graphicx}
7: \usepackage{amssymb}
8: \usepackage{latexsym}
9: \usepackage[dvips]{epsfig}
10: \renewcommand\theequation{\thesection.\arabic{equation}}
11: %\newcommand{\ip}[2]{\langle #1 | #2 \rangle}
12: \newcommand{\ipp}[3]{\langle #1 | #2 | #3 \rangle}
13: \newcommand{\derv}[2]{ \frac{d #1}{d #2}}
14: \newcommand{\dervs}[2]{ \frac{d^2 #1}{d #2^2}}
15: \newcommand{\pdrv}[2]{ \frac{\partial #1}{\partial #2}}
16: \newcommand{\pdrvs}[2]{ \frac{\partial^2 #1}{\partial #2^2}}
17: \newcommand{\unit}[1]{ \, {\rm #1} }
18: \newcommand{\half}{ \frac{1}{2}}
19: \newcommand{\dff}[1]{ {\rm d} #1 }
20: \newcommand{\vx}{\vec{x}}
21: \newcommand{\pa}{\partial}
22: \newcommand{\vp}{\varphi}
23: \newcommand{\ket}[1]{| #1 \rangle}
24: \newcommand{\bra}[1]{\langle #1 |}
25: \newcommand{\gi}{\rightarrow \infty}
26: \newcommand{\be}{\begin{equation}}
27: \newcommand{\ee}{\end{equation}}
28: \newcommand{\bea}{\begin{eqnarray*}}
29: \newcommand{\eea}{\end{eqnarray*}}
30: \newcommand{\bean}{\begin{eqnarray}}
31: \newcommand{\eean}{\end{eqnarray}}
32: \newcommand{\overleftrightarrow}[1]{\vbox{\ialign{##\crcr
33:     $\leftrightarrow$\crcr\noalign{\kern-1pt\nointerlineskip}
34:     $\hfil\displaystyle{#1}\hfil$\crcr}}}
35: \newcommand{\up}[3]{\stackrel{#1}{#2}\hspace{-3pt}^{#3}}
36: \newcommand{\rf}[1]{\mbox{(\ref{#1})} }
37: \newcommand{\ct}[1]{\mbox{(\cite{#1})} }
38: \newcommand{\ttt}[1]{\texttt{#1}}
39: \newcommand{\td}[1]{\tilde{#1}}
40: 
41: \newcommand{\n}[1]{\label{#1}}
42: \newcommand{\ind}[1]{{\mbox{\scriptsize #1}}}
43: 
44: 
45: 
46: 
47: 
48: \begin{document}
49: 
50: 
51: \title{Scattering of a Long Cosmic String by a Rotating Black Hole}
52: \author{Martin Snajdr\footnote{e-mail: msnajdr@phys.ualberta.ca} ${}^1$, 
53: Valeri Frolov\footnote{e-mail: frolov@phys.ualberta.ca} ${}^1$ and 
54: Jean-Pierre De Villiers\footnote{e-mail: jd5v@vanessa.astro.virginia.edu} ${}^2$}
55: \maketitle
56: 
57: \begin{center}
58: \noindent{
59: $^{1}${\em
60: Theoretical Physics Institute, Department of Physics, \ University of
61: Alberta, \\ Edmonton, Canada T6G 2J1}
62: }
63: 
64: \noindent  {
65: $^{2}${\em
66: Department of Astronomy, University of Virginia,
67: P.O.Box 3818,\\ University Station, Charlottesville, VA 22903-0818}
68: }
69: \end{center}
70: \bigskip
71: 
72: 
73: \maketitle
74: 
75: 
76: \begin{abstract}   
77: \noindent  
78: The scattering of a straight, infinitely long string by a rotating
79: black hole is considered. We assume that a string is moving with
80: velocity $v$ and that initially the  string is parallel to the axis
81: of rotation of the black hole. We demonstrate that as a result of
82: scattering, the string is displaced  in the direction perpendicular
83: to the velocity by an amount  $\kappa(v,b)$, where $b$ is the
84: impact parameter. The late-time solution is represented by a kink and
85: anti-kink, propagating in opposite directions at the speed of light,
86: and leaving behind them the string in a new ``phase''.  We present
87: the results of the numerical study of the string scattering and their
88: comparison with  the weak-field approximation, valid where the impact
89: parameter  is large, $b/M \gg 1$, and also with the scattering by a 
90: non-rotating black hole which was studied in earlier works.  
91: \end{abstract}
92: 
93: 
94: 
95: %\bigskip
96: \vspace{3cm}
97: 
98: 
99: \newpage
100: \section{Introduction}
101: \setcounter{equation}0
102: 
103: Study of cosmic strings and other topological defects and their
104: motion in an external gravitational field is an interesting problem.
105: Cosmic strings are topologically stable one-dimensional objects
106: which are predicted by unified theories. Cosmic strings (as well as
107: other topological defects) may appear during a phase  transition in
108: the early Universe. A detailed discussion of cosmic  strings and
109: other topological defects can be found in the book by Shellard and
110: Vilenkin \cite{ShVi:94}. Cosmic strings are naturally predicted by
111: many realistic models of particle physics. The formation of cosmic
112: strings helps to successfully exit the inflationary era in a number
113: of the inflation models motivated by particle physics
114: \cite{KoLi:87,LiRi:97}. The formation of cosmic strings is also
115: predicted in the most classes of superstring compactification
116: involving the spontaneous breaking of a pseudo-anomalous $U(1)$ gauge
117: symmetry ( see e.g. \cite{BiDePe:98} and references therein).
118: Recent measurements of CMB anisotropy and especially the position of the
119: acoustic peaks exclude some of the earlier proposed scenarios where
120: the cosmic strings are the main origin of CMB fluctuations. On the
121: other hand, recent analysis shows that a mixture of inflation and
122: topological defects is consistent with current CMB data
123: \cite{PoVa:99,Cant:00,SiPe:01,LaSh:02,BoPeRiSa:02}.
124: 
125: 
126: A black hole interacting with a
127: cosmic string is a quite rear example of interaction of two
128: relativistic non-local gravitating systems which allows rather
129: complete analysis. This makes this system interesting from pure
130: theoretical point of view. From more "pragmatic" viewpoint, this
131: system might be a strong source of gravitational waves. But in order
132: to be able to study this effect one needs first to obtain an information
133: on the motion of the string in the black hole background. The
134: situation here is very similar to the case of gravitational radiation
135: from bodies falling into the black hole.
136: 
137: In this paper we consider the scattering of a long cosmic string by a
138: rotating black hole. 
139: We assume that a string is thin and its
140: gravitational back reaction can be neglected. If $\eta$ is a
141: characteristic energy scale of a phase transition responsible for a
142: string formation then the thickness of a string is $\rho \sim
143: \eta^{-1}$ while its dimensionless mass per  unit length parameter
144: $\mu^*=G\mu/c^2 \sim \eta^{2}$. For example, for GUT strings
145: $\rho_{GUT}\approx 10^{-29}$cm and $\mu^*_{GUT}\approx 10^{-6}$, 
146: and for the electroweak phase transitions 
147: $\rho_{EW}\approx 10^{-14}$cm and $\mu^*_{EW}\approx 10^{-34}$.
148: Since $\mu^*\ll 1$ one can neglect (at least in the lowest order
149: approximation) effects connected with gravitational waves radiation
150: during the scattering of the string by a black hole. 
151: 
152: We assume that the size of the string is much larger then the
153: Schwarzschild gravitational radius $r_S=2GM/c^2$ of a black hole and
154: its total mass is much smaller than the black hole mass. The latter
155: condition together with $\mu^*\ll 1$ means that we can consider a
156: string as a test object. In order to specify the scattering problem we
157: consider the simplest setup, with the string initially far
158: away from the black hole, so that the string has the form of a straight line.
159:  We assume that
160: string is initially moving with the velocity ${\bf v}$ towards the
161: black hole and lying in the plane parallel to the rotation axis of 
162: the black hole.
163: This plane is located at distance $b$ with respect to the parallel ``plane''
164: passing through the rotation axis.
165: In analogy with a particle
166: scattering we call $b$ the {\em impact parameter}.
167: 
168: The scattering of a cosmic string by a {\em non-rotating} black hole was
169: previously studied both numerically and analytically
170: \cite{DVFr:97,DVFr:98,Page:98,DVFr:99,Page:99}.
171:  For $b\gg r_S$ the
172: string moves in the region where the gravitational potential $GM/r$ is
173: always small, and one can use a weak-field approximation where the string
174: equation of motion can be solved analytically \cite{DVFr:98,Page:98,Page:99}.
175: 
176: String scattering in the weak field limit can be described in qualitative terms.
177: In the reference frame of the string the gravitational field
178: of the black hole is time dependent and it excites the string's
179: transversal degrees of freedom. This effect occurs mainly when the
180: central part of the string passes close to the black hole. Since
181: information is propagating  along the string at the velocity of
182: light, there always exist two distant regions, `right' and `left',
183: which  have not yet felt this excitation. This asymptotic regions of the
184: string continue their motion in the initial plane (`old phase').  After
185: scattering, when string is moving again far from the black hole,  its
186: central part moves again in a plane which is parallel to the
187: initial plane but is shifted below it by a distance
188: $\kappa=2\pi GM v/\sqrt{c^2-v^2}$ (`new
189: phase'). There exist two symmetric kink-like regions separating the
190: `new' phase from the `old' one. These kinks move out of the central region
191: with the velocity of light and preserve their form. Besides the amplitude
192: $\kappa$, the kinks are characterized by a width $w$ which depends on
193: the impact parameter and the initial velocity of the string.
194: 
195: In the strong-field limit this picture qualitatively
196: remains the same for string scattering by a
197: Schwarzschild black hole. This problem was
198: analyzed in detail  numerically in  \cite{DVFr:98}. In particular, the
199: quantities $\kappa$ and $w$ were found for different values of $b$
200: and $v$. The numerical calculations in the strong field limit
201: demonstrated also that for special values of the initial impact
202: parameter $b<b_{crit}(v)$ a cosmic string is captured by the black
203: hole. The critical impact parameter for capture by the
204: Schwarzschild black hole was obtained in \cite{DVFr:97,DVFr:99}.
205: 
206: In this paper we study of scattering of straight cosmic strings by a
207: {\em rotating} black hole. We assume that the initial direction of the
208: string is parallel to the rotational axis of the
209: black hole. Our focus is on the calculation of the same
210: quantities, $\kappa$ and  $w$, and to study
211: how black hole rotation modifies the earlier results for a
212: non-rotating black hole. We present here results for string
213: spacetime evolution, real profiles of a string at different external
214: observer times $T$, as well as asymptotic scattering data. Results
215: connected with capture of the string and near-critical scattering will
216: be discussed in another paper.
217: 
218: 
219: The paper is organized as follows. Equations for string motion in a
220: curved spacetime  and their solutions in the weak field
221: regime are discussed in section~2. Section ~3 contains the
222: formulation of the initial and boundary conditions for straight
223: string motion in the Kerr geometry, and the general scheme of the numerical
224: calculations. In section~4 we describe results for strong field
225: scattering of strings in the Kerr spacetime and ``real time'' profiles
226: of the strings. Section~5 contains information about asymptotic
227: scattering data (displacement parameter and kink profiles) and their
228: dependence on the impact parameter, velocity, and angular velocity of
229: the black hole. Section~6 contains a general discussion. Numerical
230: details are supplied in an appendix.
231: 
232: 
233: \section{Equations of motion}
234: \setcounter{equation}0
235: 
236: 
237: \subsection{String equation of motion}
238: 
239: 
240: The worldsheet swept by the string in (4D curved background)
241: space-time can be parametrized by a pair of variables $\zeta^A$
242: ($A=0,1$) so that string motion is described by the equation
243: $x^{\mu}={\cal X}^{\mu}(\zeta^A)$. The dynamics of the string in a
244: spacetime with metric $g_{\mu\nu}$ is
245: described by the Nambu-Goto effective action
246: \be\n{2.1}
247: I_0[{\cal X}^{\mu}] = -\mu \int d^2\zeta\ \sqrt{-\gamma}\ ,
248: \ee 
249: where 
250: \be\n{2.2}
251: \gamma_{AB}=g_{\mu\nu}{\cal X}^\mu_{,A}\, {\cal X}^\nu_{,B}
252: \ee 
253: is the induced metric on the worldsheet.
254: The same equations of motion of the string can be derived from  the
255: Polyakov's form of the action  \cite{Poly:81},
256: \be\n{2.3} 
257: I[{\cal X}^{\mu},h_{AB}]=-{\mu \over 2}\,\int d^2\zeta
258: \sqrt{-h}h^{AB}\gamma_{AB}\, .
259: \ee
260: We use units in which $G=c=1$, and the sign conventions of \cite{MTW}. In 
261: (\ref{2.3}) $h_{AB}$ is the internal metric
262: with determinant $h$.
263: 
264: 
265: The variation of the action (\ref{2.3}) with respect to ${\cal X}^{\mu}$
266: and $h_{AB}$ gives the following equations of motion:
267: \be\n{2.5} 
268: \Box {\cal X}^{\mu}+h^{AB}{\Gamma^{\mu}}_{\alpha\beta}{\cal
269: X}^{\alpha}_{,A}{\cal X}^{\beta}_{,B}=0\, ,
270: \ee
271: \be\n{2.6} 
272: \gamma_{AB}-{1\over 2}h_{AB}h^{CD}\gamma_{CD}=0 \, ,
273: \ee
274: where
275: \be\n{2.7} 
276: \Box ={1\over \sqrt{-h}}\partial_A(\sqrt{-h}h^{AB}\partial_B)\, .
277: \ee
278: The first of these equations is the dynamical equation for string
279: motion, while the second one plays the role of a constraint. 
280: 
281: We fix internal coordinate freedom by using the gauge in which
282:    $h_{AB}$ is conformal to the flat two-dimensional metric
283:    $\eta_{AB}=\mbox{diag}(-1,1)$.
284: In this gauge
285: the equations of motion for the string have the form 
286: \be\n{2.8}
287:  \Box {\cal X}^\mu + {\Gamma^\mu}_{\alpha\beta}\, {\cal
288:  X}^{\alpha}_{,A}\, {\cal X}^{\beta}_{,B}\, \eta^{AB}\, =0\, ,
289: \ee 
290: and the constraint equations are
291: \be\n{2.9}
292: \gamma_{01} = g_{\mu\nu}\pdrv{{\cal X}^\mu}{\tau}\pdrv{{\cal X}^\nu}{\sigma} = 0\ 
293: ,
294: \ee
295: \be\n{2.10}
296: \gamma_{00}+\gamma_{11} = g_{\mu\nu}\left(\pdrv{{\cal X}^\mu}{\tau}
297: \pdrv{{\cal X}^\nu}{\tau}+\pdrv{{\cal X}^\mu}{\sigma}\pdrv{{\cal X}^\nu}{\sigma}
298: \right)\ = 0\ .
299: \ee
300: Here, $\tau\equiv\zeta^0$, $\sigma\equiv\zeta^1$ and 
301: $\Box=-\pa^2_\tau+\pa^2_\sigma$.
302: These constraints
303: are to be satisfied for the initial data. As a consequence of the
304: dynamical equations they are valid for any later time.
305: 
306: 
307: \subsection{Weak field approximation}
308: 
309: In the absence of the external gravitational field
310: $g_{\mu\nu}=\eta_{\mu\nu}$, where $\eta_{\mu\nu}$ is the flat spacetime
311: metric. In Cartesian coordinates ($T,X,Y,Z$),
312: $\eta_{\mu\nu}=\mbox{diag}(-1,1,1,1)$ and
313: ${\Gamma^{\mu}}_{\alpha\beta}=0$, and it is easy to verify that
314: \be\n{2.11} 
315: {\cal X}^{\mu}={\cal X}_0^{\mu}(\tau,\sigma)\equiv (\tau\cosh\beta,
316: \tau\sinh\beta+X_0, b,\sigma) \, ,
317: \ee
318: \be\n{2.12} 
319: h_{AB}=\eta_{AB}\equiv \mbox{diag}(-1,1)\, ,
320: \ee
321: is a solution of equations (\ref{2.5}) and (\ref{2.6}). This
322: solution describes a straight string oriented along the $Z$-axis
323: which moves in the $X$-direction with constant velocity $v=\tanh
324: \beta$. Initially, at ${\tau}_{0} = 0$, the string is found at 
325: ${\cal X}^{\mu}(0,\sigma) = (0,X_0, b,\sigma)$. Later when we use
326: this solution as initial data for a string scattering by a black
327: hole,   $b$  will play the role of impact parameter. For definiteness
328: we choose $b>0$ and $X_0<0$, so that $\beta>0$.
329: 
330: Let us consider how this solution is modified when the straight string
331: is moving in a weak gravitational field. We assume 
332: \be\n{2.13}
333: g_{\mu\nu}=\eta_{\mu\nu}+q_{\mu\nu}\, ,
334: \ee
335: \be\n{2.14}
336: {\cal X}^{\mu}(\zeta)={\cal X}_0^{\mu}(\zeta)+\chi^{\mu}(\zeta)\, ,
337: \ee
338: where $q_{\mu\nu}$ is the metric perturbation and $\chi^{\mu}$ is the
339: string perturbation. By making the perturbation of the equation of
340: motion one gets
341: \be\n{2.15}
342: \Box \chi^{\mu}=f^{\mu}\, ,
343: \ee
344: where
345: \be\n{2.16}
346: f^{\mu}=f^{\mu}(\zeta)=-{\Gamma^{\mu}}_{\alpha\beta}({\cal X}_{0})\, {\cal
347: X}_{0,A}^{\alpha}\, {\cal X}_{0,B}^{\beta}\, \eta^{AB}\, .
348: \ee
349: We use Cartesian coordinates here so that ${\Gamma^{\mu}}_{\alpha\beta}$
350: is simply the Christoffel symbol for $q_{\mu\nu}$.
351: 
352: The linearized constraint equations are
353: \be\n{2.17}
354: \eta_{\mu\nu}\pdrv{{\cal X}^\mu_0}{\tau}\pdrv{\chi^\nu}{\sigma}+
355: \eta_{\mu\nu}\pdrv{\chi^\mu}{\tau}\pdrv{{\cal X}^\nu_0}{\sigma} +
356: q_{\mu\nu}\pdrv{{\cal X}^\mu_0}{\tau}\pdrv{{\cal X}^\nu_0}{\sigma} = 0\ ,
357: \ee
358: \be\n{2.18}
359: 2\eta_{\mu\nu}\left(\pdrv{{\cal X}^\mu_0}{\tau}\pdrv{\chi^\nu}{\tau}+
360: \pdrv{{\cal X}^\mu_0}{\sigma}\pdrv{\chi^\nu}{\sigma} \right)+
361: q_{\mu\nu}\left(\pdrv{{\cal X}^\mu_0}{\tau}\pdrv{{\cal X}^\nu_0}{\tau}+
362: \pdrv{{\cal X}^\mu_0}{\sigma}\pdrv{{\cal X}^\nu_0}{\sigma} \right) = 0\ .
363: \ee
364: As in the exact non-linear case, if these linearized constraints
365: are satisfied at the initial moment of time $\tau$ they are
366: also valid for any $\tau$ for a solution $\chi^{\mu}$ of the dynamical
367: equations (\ref{2.15}).
368: 
369: 
370: The linearized equations can be used to study string motion
371: in the case where it is far away from the black hole. In this case the
372: gravitational field can be approximated as follows:
373: \be\n{2.19}
374:  ds^2  = -(1-\frac{2M}{R})dT^2+(1+\frac{2M}{R})\, (dX^2+dY^2+dZ^2) 
375:            -\frac{4J}{R^3}(XdY-YdX)dT\ ,
376: \ee
377: where $R^2=X^2+Y^2+Z^2$, and $M$ and $J=aM$ are the mass and angular
378: momentum of the black hole, respectively. In fact this asymptotic
379: form of the metric is  valid for any arbitrary stationary localized
380: distribution of matter, provided that the observer is located far from
381: it. In agreement with   (\ref{2.19}) the gravitational field
382: perturbation can be presented as
383: \be\n{2.20}
384: q_{\mu\nu}=q_{\mu\nu}^{N}+q_{\mu\nu}^{LT}\, ,
385: \ee
386: where the Newtonian and Lense-Thirring parts are
387: \be\n{2.21}
388: q_{\mu\nu}^{N}=2\varphi \delta_{\mu\nu}\, ,\hspace{1cm}
389: \varphi={M\over R}\, , \hspace{1cm}  q_{\mu\nu}^{LT}={4J\over R^3}\,
390: \delta^0_{(\mu}\, \epsilon_{\nu)\alpha 0 3} X^{\alpha}\, .
391: \ee
392: Here $\epsilon_{\alpha\beta\gamma\delta}$ is the antisymmetric symbol.
393: The Lense-Thirring part $q_{\mu\nu}^{LT}$ of the metric \cite{MTW,ThLe:18} is produced
394: by the rotation of the source of the gravitational  field and it is
395: responsible for frame dragging.
396: 
397: \subsection{Newtonian scattering}
398: 
399: In the linear approximation we can study the action of each of the 
400: parts of the metric perturbations  independently. For the Newtonian
401: part one has the following expression for the force $f^{\mu}_{N}$
402: \bean\n{2.22}
403: f^{0}_{N} &=&2\sinh(\beta)\cosh(\beta)\, ,\varphi_{,1}\\
404: f^{1}_{N} &=& 0\, , \\
405: f^{2}_{N} &=& -2\sinh^2\beta\varphi_{,2}\, ,\\
406: f^{3}_{N} &=& -2\cosh^2\beta\varphi_{,3}\ ,
407: \eean
408: and the  constraint equations read
409: \bean\n{2.26}
410: \chi^3_{,\tau}-\cosh\beta\chi^0_{,\sigma}+
411: \sinh\beta\chi^1_{,\sigma} &=& 0 \, ,\\
412: \chi^3_{,\sigma}-\cosh\beta\chi^0_{,\tau}+
413: \sinh\beta\chi^1_{,\tau} &=& -2\varphi\cosh^2\beta\ .
414: \eean
415: 
416: The result $f^1=0$ means that in the first order
417: approximation the string lies in the $YZ-$plane at any fixed time
418: $\tau=\mbox{const}$. The dynamical equations (\ref{2.3})
419: can be easily integrated by using the retarded Green function (see
420: \cite{DVFr:98}).
421: 
422:  We focus now our attention on the behavior of
423: $\chi^2$ which describes the string deflection in the plane
424: orthogonal to its motion. It is easy to show that the asymptotic value
425: of $\chi^2(\tau\to\infty,\sigma)$ is the same for any fixed
426: $\sigma$. We denote it $b-\kappa$. The following expression is valid
427: for $\kappa$
428: \be\n{2.28}
429: \kappa=-{1\over 2 \sinh \beta}\, \int_{\Pi_0}\, dX \, dZ f^{2}\, ,
430: \ee 
431: where $\Pi_0$ is the  two-dimensional worldsheet swept by the
432: string in its motion in the background spacetime.
433: This integral can be easily calculated since 
434: \be\n{2.29}
435: \int_{\Pi_0}\, dX \, dZ \varphi_{,Y}=2\pi M\, 
436: \ee 
437: is the flux of the Newtonian field strength through $\Pi_0$. Thus we
438: have
439: \be\n{2.30}
440: \kappa=2\pi M \sinh\beta\, .
441: \ee
442: 
443: At any late moment of time $\tau$ the central part of the string is a
444: straight line moving in the plane $Y=b-\kappa$, while its far distant
445: parts move in the original plane $Y=b$. These different ``phases'' are
446: connected by kinks propagating in the direction from the center with
447: the velocity of light. (For details, see \cite{DVFr:99}). We call
448: $\kappa$ the {\em displacement parameter}.
449: Figure~\ref{cs_Y} shows the Y-direction displacement for the Newtonian 
450: scattering as a function of $\tau$ and $\sigma$.
451: 
452: 
453: \begin{figure}[ht]
454: \begin{center}
455: \epsfig{file=newton_tsy.eps, width=10cm}
456: \caption{$Y-$profile for Newtonian scattering}
457: \label{cs_Y}
458: \end{center}
459: \end{figure}
460: 
461: 
462: \subsection{Lense-Thirring scattering }
463: 
464: 
465: 
466: 
467: In the presence of rotation the Lense-Thirring force acting on the
468: string in the linearized approximation is
469: \bean\n{2.31}
470: f^{0}_{LT}  &=& 6J\sinh^2\beta\frac{b (X_0+\tau\sinh\beta)}{{\cal
471: R}^5} \, ,\\
472: \n{2.32}
473: f^{1}_{LT}   &=& 0 \, ,\\
474: \n{2.33}
475: f^{2}_{LT}   &=& 2J\sinh\beta\cosh\beta\left(\frac{1}{{\cal R}^3}-
476: \frac{3\sigma^2}{{\cal R}^5}\right) \, ,\\
477: \n{2.34}
478: f^{3}_{LT}   &=& 6J\sinh\beta\cosh\beta \frac{b\sigma}{{\cal R}^5}\, ,
479: \eean
480: where ${\cal R}=\sqrt{(X_0+\tau\sinh\beta)^2+b^2+\sigma^2}$.
481: As before, $f^1=0$.
482: 
483: The constraint equations now take the form
484: \be\n{2.35}
485: \chi^3_{,\tau} - \cosh\beta\chi^0_{,\sigma}+\sinh\beta\chi^1_{,\sigma}
486: = 0\, ,
487: \ee
488: \be\n{2.36}
489: \chi^3_{,\sigma} - \cosh\beta\chi^0_{,\tau}+\sinh\beta\chi^1_{,\tau} 
490: = -2J\sinh\beta\cosh\beta\frac{b}{{\cal R}^3}\, .
491: \ee
492: 
493: 
494: 
495: The dynamical equations can be  solved analytically.
496: For the initial conditions
497: \be\n{2.37}
498: \chi^0|_{\tau=0} = 0\, ,
499: \ee
500: \be\n{2.38}
501: \chi^2|_{\tau=0} =\pdrv{\chi^2}{\tau}|_{\tau=0} = 0\, ,
502: \ee
503: \be\n{2.39}
504: \chi^3|_{\tau=0} =\pdrv{\chi^3}{\tau}|_{\tau=0} =0\, ,
505: \ee
506: the solution is
507: \[
508: \chi^0 = Jb\sinh\beta\left(\frac{(X_0+\tau\sinh\beta)\sinh\beta
509: +\sigma}{\Delta_{(+-)}{\cal R}}
510: + \frac{(X_0+\tau\sinh\beta)\sinh\beta-\sigma}{\Delta_{(++)}{\cal
511: R}}\right. \, 
512: \]
513: \be\n{2.40}
514: - \left.\frac{X_0\sinh\beta-(\tau-\sigma)}{\Delta_{(+-)}\rho_-}
515: - \frac{X_0\sinh\beta-(\tau+\sigma)}{\Delta_{(++)}\rho_+}\right) \, ,
516: \ee
517: \be\n{2.41}
518: \chi^1 = 0\, ,
519: \ee
520: \[
521: \chi^2 = J\sinh\beta\cosh\beta
522: \left(-\frac{\tau^2\sinh^2\beta+\tau\sigma\sinh^2\beta
523: +2X_0\tau\sinh\beta+\sigma X_0\sinh\beta+X_0^2  +
524: b^2}{\Delta_{(++)}{\cal R}}\right. 
525: \]
526: \be\n{2.42}
527: +\frac{-\tau^2\sinh^2\beta+\tau\sigma\sinh^2\beta
528: -2X_0\tau\sinh\beta+\sigma X_0\sinh\beta-X_0^2 -
529: b^2}{\Delta_{(--)}{\cal R}}
530: \ee
531: \[
532: + \left.\frac{X_0\sinh\beta(\tau+\sigma)+X_0^2
533: +b^2}{\Delta_{(++)}\rho_+} +\frac{X_0\sinh\beta(\tau-\sigma)+X_0^2
534: +b^2}{\Delta_{(--)}\rho_-}\right)\, ,
535: \]
536: \[
537: \chi^3 = Jb\sinh\beta\cosh\beta\left(\frac{(X_0+\tau\sinh\beta)\sinh\beta
538: -\sigma}{\Delta_{(++)}{\cal R}}
539: -\frac{(X_0+\tau\sinh\beta)\sinh\beta+\sigma}{\Delta_{(--)}{\cal R}}\right.
540: \]
541: \be\n{2.43}
542: - \left.\frac{X_0\sinh\beta-(\tau+\sigma)}{\Delta_{(++)}\rho_+} 
543: +\frac{X_0\sinh\beta-(\tau-\sigma)}{\Delta_{(--)}\rho_-}\right)\, ,
544: \ee
545: \be\n{2.44}
546: \Delta_{\pm\pm} = (X_0\pm(\tau\pm\sigma)\sinh\beta)^2 +
547: b^2\cosh^2\beta\, ,
548: \ee
549: \be
550: \rho_{\pm} = \sqrt{X_0^2+b^2+(\tau\pm\sigma)^2}\, .
551: \ee
552: 
553: Using these relations it is possible to see that the displacement
554: parameter $\kappa=\lim_{\tau\to\infty}\, \chi^2(\tau,\sigma=0)$
555: vanishes for this solution. The reason  is the following. Using
556: (\ref{2.33}) it is easy to present $f_{LT}^2$ as
557: \be\n{2.44a}
558: f_{LT}^2=2J\sinh\beta\, \cosh\beta {\partial\over \partial
559: \sigma}\left({\sigma\over {\cal R}^3}\right)\, .
560: \ee
561: The asymptotic displacement is given by (\ref{2.28}). The integral
562: over $Z$ or, equivalently, the integral over $\sigma$ from
563: $f_{LT}^2$, which is a total derivative over $\sigma$, reduces to
564: boundary terms $\sigma/{\cal R}^3$ at $\sigma=\pm\infty$, which
565: vanish.
566: 
567: 	
568: 
569: \begin{figure}[ht]
570: \begin{center}
571: \epsfig{file=lt_tsy.eps, width=10cm}
572: \caption{$Y-$profile for Lense-Thirring scattering}
573: \label{cs_lt}
574: \end{center}
575: \end{figure}
576: 
577: Figure~\ref{cs_lt} shows the $Y-$direction displacement as a
578: function of $\tau$ and $\sigma$ for the weak field scattering. It
579: should be emphasized that the scale of structures for
580: Lense-Thirring scattering is much smaller than that for
581: Newtonian scattering. This can be easily seen if we compare the
582: Newtonian force $f_N\sim M/{\cal R}^2$ with the Lense-Thirring force
583: $f_{LT}\sim J/{\cal R}^3$
584: \be
585: {f_{LT}\over f_N}\sim {J\over M{\cal R}}\le {J\over Mb}\, .
586: \ee
587: For the scattering by a rotating black hole $J=aM$, where $|a|/M \le
588: 1$ is the rotation parameter. Hence
589: \be
590: {f_{LT}\over f_N} \le {M\over b}\, 
591: \ee
592: which is small for the weak field scattering. For this reason, in the
593: weak field regime the string profiles for the prograde or
594: retrograde  scattering by rotating black hole do not greatly
595: differ from the profiles for the scattering by a non-rotating black
596: hole of the same mass. The situation is 
597: different for strong-field scattering: the displacement parameter
598: is different for prograde and retrograde scattering, and for near-critical
599: scattering the profiles of the kinks contain visible
600: structure produced by effects connected with the rotation of the
601: black hole.
602: 
603: 
604: \section{String scattering by a Kerr black hole}
605: \setcounter{equation}0
606: 
607: \subsection{String in the Kerr spacetime}
608: 
609: Our aim is to study a string motion in the Kerr spacetime. The Kerr 
610: metric in  Boyer-Lindquist coordinates $(t,r,\theta,\phi)$ has 
611: the form
612: \bean\n{3.1}
613: ds^2 &=&  -(1-\frac{2Mr}{\Sigma})dt^2+\frac{\Sigma}{\Delta}dr^2 +
614:          \Sigma d\theta^2 + \frac{A\sin^2{\theta}}{\Sigma}d\phi^2 -
615:          \frac{4aMr\sin^2{\theta}}{\Sigma}dtd\phi \ ,\\
616: \Sigma &=& r^2+a^2\cos^2{\theta} \ ,\\
617: \Delta &=& r^2 - 2Mr + a^2 \ ,\\
618: A &=& (r^2+a^2)^2 - a^2\Delta \sin^2{\theta} \ ,
619: \eean
620: where $M$ is the mass of the black hole, and $J=aM$ is its angular momentum 
621: ($0\le a\le M$). At far distances this metric has the asymptotic form
622: (\ref{2.19}).
623: 
624: 
625: 
626: 
627: In order to be able to deal with the case where part of the string
628: crosses the event horizon for the numerical simulation we adopt the
629: so-called Kerr  (in-going) coordinates $(\td{v},r,\theta,\td{\phi})$
630: \[
631: ds^2 = -(1-\frac{2Mr}{\Sigma})d\td{v}^2 + 2d\td{v}dr - 
632:          \frac{4aMr\sin^2{\theta}}{\Sigma}d\td{v}d\td{\phi} -
633:          a\sin^2{\theta}drd\td{\phi}
634: \]	 
635: \be \n{3.5}
636: + \Sigma d\theta^2 + \frac{A\sin^2{\theta}}{\Sigma}d\td{\phi}^2\ ,
637: \ee
638: with
639: \bean\n{3.6}
640: \td{v} &=& t+r+M\ln|{\Delta\over 4M^2}| + \frac{M^2}{\sqrt{M^2-a^2}}\ln
641:       \left|\frac{r-M-\sqrt{M^2-a^2}}{r-M+\sqrt{M^2-a^2}}\right| + M\ ,\\
642: \td{\phi} &=& \phi+\frac{a}{2\sqrt{M^2-a^2}}\ln
643:          \left|\frac{r-M-\sqrt{M^2-a^2}}{r-M+\sqrt{M^2-a^2}}\right|\ .
644: \eean
645: The metric (\ref{3.1}) has the asymptotic form
646: \be\n{3.7}
647: ds^2 = -dt^2 + dr^2 + r^2(d\theta^2 + \sin^2\theta d\phi^2) + 
648:        \frac{2M}{r}(dt^2+dr^2) - \frac{4aM}{r}\sin^2\theta dtd\phi\ .
649: \ee
650: Let us introduce new ``quasi-Cartesian'' coordinates
651: \bea\n{3.8}
652: T &=& t\ ,\\
653: X &=& R\sin{\theta}\cos{\phi}\ ,\\
654: Y &=& R\sin{\theta}\sin{\phi}\ ,\\
655: Z &=& R\cos{\theta}\ , 
656: \eea
657: where 
658: \be\n{3.8b}
659: r=R+M\ .
660: \ee
661: One can easily check that the metric (\ref{3.7}) in the quasi-Cartesian coordinates (\ref{3.8})
662: has the asymptotic form (\ref{2.19})
663: 
664: It should be emphasized that without the shift (\ref{3.8b}) of the radial coordinate one does
665: not recover (\ref{2.19}). The reason for this is easy to understand if one considers the same
666: problem for the Schwarzschild geometry. The asymptotic limit of (\ref{2.19}) with $a=0$ can 
667: be found in the isotropic coordinates with $r = \tilde{R}(1+M/2\tilde{R})^2$ in which the 
668: spatial part of the metric is conformal to the flat metric. Asymptotically, it is sufficient
669: to use (\ref{3.8b}) which is the leading part of $r$ at large $\tilde{R}$.
670: 
671: In what follows we shall use these quasi-Cartesian coordinates
672: for representing the position and the form of the string in the Kerr
673: spacetime even if we are not working in the weak field regime.
674: It should be emphasized that since the space in the Kerr
675: geometry is not flat, plots constructed in these coordinates do not
676: give a ``real picture''. This is a special case of a general problem
677: of the visualization of physics in a 4-dimensional curved spacetime.
678: 
679: 
680: 
681: \subsection{Initial and boundary conditions}
682: \label{sec3.2}
683: 
684: In an ideal situation, we would study scattering by initially placing
685: our infinitely long string at spatial infinity,
686: where the spacetime is flat and the straight string can be described in
687: simple terms.
688: In a numerical scheme, however, the string cannot be infinitely long
689: and we must start the simulation at a finite distance  from the black
690: hole. We discuss here the initial and boundary conditions used for the
691: simulations.
692: 
693: In studying the scattering of a straight string, we
694: consider the special case where the string is initially parallel to the
695: axis of rotation of the black hole. We use Cartesian coordinates
696: in the asymptotic region so that $X$-axis  coincides with the
697: direction of motion and the $Z$-axis is along the string
698: (see Fig.~\ref{InitSetup}).
699: 
700: \begin{figure}[ht]
701: \begin{center}
702: \epsfig{file=fig01.eps, width=7cm}
703: \caption{Initial setup of the scattering experiment}
704: \label{InitSetup}
705: \end{center}
706: \end{figure}
707: 
708: 
709: 
710: 
711: 
712: We consider a string segment of length $L\sim 10^4M$ at a
713: time $\tau_0$ and an
714: initial distance $X_0$ from the black hole.
715: In order to keep accuracy high and yet prevent the calculation
716: time from being inordinately large, we do not evolve the straight string
717: numerically from this initial position. Instead, we
718: use the weak-field approximation to describe the string
719: configuration at a later time
720: $\tau_s$ where the distance $X_s$ is closer to the black hole, $|X_s|\ll |X_0|$
721: ~\footnote{ In our simulations we
722: put $X_0 = -10^6 M$ and $X_s = -500 M$  }.
723: 
724: Given a sufficiently long string segment, the boundary points move
725: at a great distance
726: from the black hole, so that their evolution can be described by the
727: weak field approximation until information about the
728: interaction with the black hole reaches the boundary. We denote this
729: time as $\tau_*$. Starting from this moment we solve the dynamical
730: equations in the region $|\sigma|< L-(\tau-\tau_*)$ (see
731: figure~\ref{f02}). The larger is the initial value $L$ the longer one can
732: go in $\tau$ in the simulation. We choose $L$ to be large enough to
733: provide the required accuracy in the determining the final scattering
734: data.
735: 
736: Since the boundary conditions found from the weak field regime calculations
737: are not exact there will be disturbance created by this effect.
738: In some cases, when it is important to exclude
739: these disturbances, we used a modified scheme of calculations in which 
740: no boundary conditions are used from the very beginning of the simulation.
741: The price for this is that one needs to take longer initial size of 
742: $L$ which in turn increases the computational time.
743: 
744: 
745: \begin{figure}[ht]
746: \begin{center}
747: \epsfig{file=fig02b.eps, width=\textwidth}
748: \caption{Scheme of  time domains for scattering problem}
749: \label{f02}
750: \end{center}
751: \end{figure}
752: 
753: 
754: \subsection{Solving dynamical equations and constraints}
755: 
756: Using the initial conditions at $\tau_s$ and boundary conditions we
757: solve numerically the dynamical equations (\ref{2.8}).
758: Since the equations are
759: of second order, we use the weak-field approximation to get initial data at
760: $\tau_s$ and $\tau_s+\Delta\tau$ in order to completely specify the initial
761: value problem.  The numerical scheme uses second-order finite differences and
762: evolves the string configuration using an implicit scheme.  Because of the
763: symmetry $\sigma\to -\sigma$ it is sufficient to evolve only half of the string
764: worldsheet, $\sigma\le 0$, and use a reflecting
765: boundary condition at the string midpoint. 
766: The numerical grid is non-uniform with the denser part following the motion
767: of the kink.
768: We used the
769: constraint equations (\ref{2.9}-\ref{2.10}) for an independent check of the
770: accuracy of the calculations.  More details on the numerical scheme
771: can be found in the Appendix.
772: 
773: 
774: \section{String profiles for strong field scattering}
775: \setcounter{equation}0
776: 
777: \subsection{General picture}
778: 
779: 
780: \begin{figure}
781: \begin{tabular}{cc}
782: \epsfig{file=scat_i80p0_a0p0_b1p0_tsk.eps, width=7cm}
783: &
784: \epsfig{file=scat_i6p0_a0p0_b1p0_tsk.eps, width=7cm}\\
785: {\bf a} & {\bf b} \\
786: \epsfig{file=scat_i6p0_a1p0_b1p0_tsk.eps, width=7cm}
787: &
788: \epsfig{file=scat_i6p0_a-1p0_b1p0_tsk.eps, width=7cm}\\
789: {\bf c} & {\bf d} \\
790: \end{tabular}
791: \caption{Displacement in $Y-$ direction as a function of
792: $(\tau,\sigma)$ for given velocity $v/c=0.762$ in the weak (a) and
793: strong (b--d) field regime.}
794: \label{3d}
795: \end{figure}
796: 
797: 
798: Figure~\ref{3d} demonstrates general features of straight string scattering,
799: here for a string with velocity $v=0.762$ ($\beta=1$).  The four panels show the
800: displacement in the $Y-$direction as a function of internal coordinates
801: $(\tau,\sigma)$ for weak and strong field scattering.
802: 
803: Figure~\ref{3d} (a) shows the scattering for an impact parameter $b=80\, M$.
804: The value of the displacement parameter at the largest $\tau$ shown is
805: $\kappa=7.23M$.  This value as well as the form of the kinks are in a very good
806: agreement with the weak-field scattering results.
807: Plots for prograde and retrograde scattering at $b=80\, M$ are
808: virtually indistinguishable from Fig.~\ref{3d} (a).
809: 
810: Figures~\ref{3d} (b--d) show the scattering for  $b=6\, M$.
811: Qualitatively these plots are similar to weak-field scattering. The main
812: differences are the following: (1) The value of the displacement parameter
813: $\kappa=11.03M$ for the non-rotating
814: black hole is significantly larger then $\kappa=7.23$ for $b=80\, M$.
815: (2) Effects of rotation are
816: more pronounced. The displacement $\kappa$ for retrograde scattering is
817: essentially larger than $\kappa$ for the prograde scattering. (3) The
818: string profile after scattering is more sharp, the width of the kinks
819: is visibly smaller that for weak-field scattering.
820: 
821: 
822: \subsection{``Real-time profiles'' of the string for strong-field scattering}
823: 
824: \begin{figure}
825: \begin{tabular}{c}
826: \epsfig{file=scat_i6p0_a0p0_b1p0_rho_XYZ.eps, width=8cm}
827: \end{tabular}
828: \hfill
829: \begin{tabular}{c}
830: \epsfig{file=scat_i6p0_a0p0_b1p0_rho_XZ.eps, width=4.5cm}\\
831: \epsfig{file=scat_i6p0_a0p0_b1p0_rho_YZ.eps, width=4.5cm}
832: \end{tabular}
833: \caption{``Real-time profiles'' of the string and their $XZ-$ and
834: $YZ-$ projection for the strong field scattering by the Schwarzschild
835:  black hole.}
836: \label{RT_S}
837: \end{figure}
838: 
839: \begin{figure}
840: \begin{tabular}{c}
841: \epsfig{file=scat_i6p0_a-1p0_b1p0_rho_XYZ.eps, width=8cm}
842: \end{tabular}
843: \hfill
844: \begin{tabular}{c}
845: \epsfig{file=scat_i6p0_a-1p0_b1p0_rho_XZ.eps, width=4.5cm}\\
846: \epsfig{file=scat_i6p0_a-1p0_b1p0_rho_YZ.eps, width=4.5cm}
847: \end{tabular}
848: \caption{``Real-time profiles'' of the string and their $XZ-$ and
849: $YZ-$ projection for the strong field retrograde scattering by the 
850: extremal Kerr black hole.}
851: \label{RT_K-}
852: \end{figure}
853: 
854: 
855: \begin{figure}
856: \begin{tabular}{c}
857: \epsfig{file=scat_i6p0_a1p0_b1p0_rho_XYZ.eps , width=8cm}
858: \end{tabular}
859: \hfill
860: \begin{tabular}{c}
861: \epsfig{file=scat_i6p0_a1p0_b1p0_rho_XZ.eps, width=4.5cm}\\
862: \epsfig{file=scat_i6p0_a1p0_b1p0_rho_YZ.eps, width=4.5cm}
863: \end{tabular}
864: \caption{``Real-time profiles'' of the string and their $XZ-$ and
865: $YZ-$ projection for the strong field prograde scattering by the 
866: extremal Kerr black hole.}
867: \label{RT_K+}
868: \end{figure}
869: 
870: It should be emphasized that figure~\ref{3d} give an accurate impression of the
871: displacement effect, but they give a distorted view of the real form of the
872: string.  The reason is evident.  The grid imposed on the worldsheet is
873: determined by the choice of $(\tau,\sigma)$ coordinates.  But a $\tau={\rm const}$
874: section differs from a time $T={\rm const}$ section in the laboratory slice of
875: spacetime.  The position of the string at given time $T$ can be found by using
876: the functions $T(\tau,\sigma)$, $X(\tau,\sigma)$, $Y(\tau,\sigma)$,
877: $Z(\tau,\sigma)$, to find functions $X(T,Z)$ and $Y(T,Z)$.  For fixed $T$ these
878: functions determine a position of the string line in 3-space.
879: 
880: 
881: Figures~\ref{RT_S}, \ref{RT_K-} and \ref{RT_K+} show a sequence of
882: ``real-time'' profiles for strong-field string scattering at five
883: different coordinate times $T$. We ordered them so that the larger
884: number labeling the string corresponds to later time.
885: In figure~\ref{RT_S} the black hole is non-rotating.
886: Figures~\ref{RT_K-} and \ref{RT_K+} show the ``real-time'' profiles
887: for retrograde and prograde scattering for a maximally rotating black
888: hole, respectively. Again we can see that due to frame dragging, the
889: distortion of the string is more pronounced for retrograde scattering.
890: 
891: 
892: 
893: \section{Late time scattering data}
894: \setcounter{equation}0
895: 
896: \subsection{Displacement parameter}
897: 
898: \begin{figure}
899: \begin{tabular}{c}
900: \hfill 
901: \epsfig{file=k_vs_i_v1p0f.eps, width=8cm}\\
902: \epsfig{file=k_vs_i_v2p0f.eps, width=8cm}\hfill
903: \epsfig{file=k_vs_i_v3p0f.eps, width=8cm}
904: \end{tabular}
905: \caption{Displacement parameter $\kappa$ as a function of the impact
906: parameter $b$ for different velocities $v$.}
907: \label{disp}
908: \end{figure}
909: 
910: Figure~\ref{disp} shows the dependence of the displacement parameter $\kappa$ on
911: the impact parameter $b$ for different velocities $v$.  For  a given value of $b$
912: the curve for $a/M=-1$ always lies higher than the curves for more positive 
913: values of
914: $a/M$.  This is true for all values of velocity $v$, but this difference is more
915: pronounced at large velocities.  
916: 
917: The relative location of the curves for given $v$ and $b$ and 
918: different $a$ can be explained by frame dragging.
919: Namely,  a string in the retrograde motion is effectively slowed down
920: when it passes near the black hole and hence it spends more time near
921: it. As the result its displacement is greater than the displacement
922: for the prograde motion with the same impact parameter. Let us make
923: order of magnitude estimation of this effect. For this
924: purpose we use metric (\ref{2.19}). We focus our attention on the
925: Lense-Thirring term and neglect for a moment the Newtonian part.
926: Consider first a point particle moving in this metric with the velocity $v$
927: and impact parameter $b$. In the absence of rotation it moves with
928: constant velocity so that $(X=vT, Y=b, Z=0)$ and the proper time $\tau$ is
929: \be
930: \tau_0 =\sqrt{1-v^2}\, T\, .
931: \ee
932: In the presence of rotation one  has
933: \be
934: d\tau^2=(1-v^2)\, dT^2 -{4Jbv\over R^3}\, dT^2\, ,
935: \ee
936: where
937: \be
938: R^2=b^2+v^2\, T^2\, .
939: \ee
940: For large $b$ one can write
941: \be
942: \tau =\tau_0 +\Delta\tau\, ,
943: \ee
944: where
945: \be
946: \Delta\tau = -{2Jbv\over \sqrt{1-v^2}}\, \int_{-\infty}^{\infty}\,
947: {dT\over (b^2+v^2T^2)^{3/2}}\, =-{4J\over b\sqrt{1-v^2}}\, .
948: \ee
949: This quantity $\Delta\tau$ characterize additional time delay for the
950: motion in the Lense-Thirring field. The characteristic time of motion
951: in the vicinity of a black hole is $\tau_{int}\sim b/v$. Thus we have
952: \be
953: {\Delta\tau\over \tau_{int}}\sim -{4Jv\over b^2\sqrt{1-v^2}}\, .
954: \ee
955: One can expect that a similar delay takes place for a string motion.
956: As a result of being longer close to the black hole the string
957: has a larger displacement by the value $\Delta \kappa$ such that
958: \be
959: {\Delta\kappa\over \kappa}\sim {\Delta\tau\over \tau_{int}}\sim
960: -{4Ma\over b^2}\sinh \beta \, . 
961: \n{5.7}
962: \ee
963: Qualitatively this relation explains dependence of $\kappa$ on $a$ and
964: $\beta$ presented in figure~\ref{disp}.
965: 
966: 
967: More generally, if we look at the $\alpha=a/M$ in the metric as a parameter, we
968: can use perturbation theory to calculate the effect of rotation.
969: The solution ${\cal X}^\mu$ can be expanded in powers of $\alpha$
970: \be
971: {\cal X}^\mu = \up{0}{\cal X}{\mu}+\alpha\up{1}{\cal X}{\mu}
972:                 +\alpha^2\up{2}{\cal X}{\mu} + \dots\ ,
973: \ee
974: where $\up{0}{\cal X}{\mu}(\tau,\sigma)$ is the solution for the 
975: non-rotating case and 
976: $\up{i}{\cal X}{\mu}(\tau,\sigma)$ are the perturbation corrections.
977: 
978: Numerical calculations show that the first two terms in the perturbation series 
979: give very good approximation to the solution as long as we are  not
980: close to critical scattering. As a consequence 
981: \be\n{5.9} 
982: \kappa = \kappa_0 - \kappa_1\alpha - \kappa_2\alpha^2\ ,
983: \ee
984: where $\kappa_0$ is the $Y$ displacement for the non-rotating black hole, and 
985: $\kappa_1$,$\kappa_2$ are defined simply as
986: \be
987: \kappa_{1,2} = \lim_{\tau\to\infty} \up{1,2}{\cal X}{2}(\tau,0)\ .
988: \ee
989: In the above $\up{1,2}{\cal X}{2}$ denotes the $Y$ coordinate corrections.
990: The equations of motion for $\up{1,2}{\cal X}{\mu}$ are linear and 
991: $\up{1,2}{\cal X}{\mu}$ can be obtained in one run together with 
992: $\up{0}{\cal X}{\mu}$.
993: 
994: In order to determine the coefficients in equation~(\ref{5.9}) we need 
995: to extrapolate the relevant data obtained during the simulation to infinity.
996: One possibility is to numerically advance the solution very far from the black hole and simply
997: take the value at the end of the simulation as our estimate. The problem with this approach is that
998: to advance very far by maintaining high accuracy of the solution
999: is a very lengthy process, feasible only for larger impact parameters (since the grid can be
1000: sparser).
1001: 
1002: Our approach was to advance to a moderate distance ($X_{\rm max}\approx 8000M$) keeping high accuracy of the
1003: solution (judged by the constraint equations).
1004: Then we fit the calculated data by the function
1005: \be
1006: \kappa(0,\tau) = \kappa + k(\tau-\tau_{sh})^\gamma\ ,
1007: \ee
1008: where $\kappa$, $k$, $\tau_{sh}$, $\gamma$ are the parameters to be fitted.
1009: We take our first data point at $X=1000M$.
1010: To see how consistent this procedure is and to estimate the errors
1011: we perform three fits with different set of data points.
1012: The first set uses all the data points,
1013: the second one uses only the first half,
1014: and the third one uses only the last half of the data points.
1015: For this particular problem the disturbances from the boundary have adverse effect
1016: to the fit therefore we use the null boundary approach.
1017: 
1018: 
1019: Our experience shows that the fitting procedure for $\kappa$ works much better 
1020: for larger velocities and larger impact parameters.
1021: For the velocities $v/c = 0.762$, $0.964$, $0.995$ the estimated values of 
1022: $\kappa$ from all three fits are consistent and differ by no more than 3\%
1023: \footnote{For $b\ge12M$ the results from the three fits differ by less than 1\%.}.
1024: This fact gives us reasonable confidence that the plotted values of $\kappa$ 
1025: indeed represent the true values.
1026: 
1027: Unfortunately the situation is not so good for slower velocities. For example for 
1028: the case $v/c=0.462$ the slope of the 
1029: fitted function $\kappa(0,\tau)$ is almost constant during the later phases
1030: of the simulation and fits using different sets of data points yield 
1031: inconsistent (and improbable) values of $\kappa$.
1032: We expect the behavior of $\kappa(0,\tau)$ to change but we did not see it in our data 
1033: even after driving the simulation four times further then for the other velocities 
1034: ($X_{\rm \max}\approx 32,000M$).
1035: Again, this behavior is more pronounced for smaller impact parameters $b$.
1036: 
1037: 
1038: \begin{figure}[!htb]
1039: \begin{tabular}{c}
1040: \epsfig{file=k1_vs_i.eps, width=8cm}
1041: \hfill
1042: \epsfig{file=k2_vs_i.eps, width=8cm}\\
1043: \end{tabular}
1044: \caption{Plots of $\kappa_1$ and $\kappa_2$ for different velocities $v$
1045: and impact parameters $b$. The straight lines represent a linear fit.}
1046: \label{kappa_12}
1047: \end{figure}
1048: \begin{figure}[!htb]
1049: \begin{center}
1050: \epsfig{file=k1overk_vs_i.eps, width=12cm}
1051: \caption{Plot of $\kappa_1/\kappa$ for different velocities $v$ 
1052: and impact parameters $b$. The straight lines represent a linear fit.}
1053: \label{k1_over_k}
1054: \end{center}
1055: \end{figure}
1056: 
1057: To obtain the values for $\kappa_1$ and $\kappa_2$ we use the same method as for
1058: $\kappa$.
1059: We did not encounter any difficulties in estimating the values of $\kappa_1$ and
1060: $\kappa_2$ for any of the four velocities.
1061: The trend here is the opposite of the one for $\kappa$ -- the fits work
1062: best for smaller velocities and smaller impact parameters.
1063: The fits with different set of data points are typically less than $1$\% appart.
1064: 
1065: Figure~\ref{kappa_12} shows the results in a log-log plot.
1066: Figure~\ref{k1_over_k} shows the plot of $\kappa_1/\kappa$.
1067: The straight lines shown are obtained from least square fit.
1068: Note that the plots for the velocities $v/c=0.964$ and $v/c=0.995$
1069: lie on top of each other.
1070: 
1071: In the interval shown $\kappa_1$ and $\kappa_2$ 
1072: can be reasonably approximated by a functions of the form
1073: \be
1074: \n{5.12}
1075: \kappa_{1,2} \approx A_{1,2}(v)\left(\frac{1}{b}\right)^{\lambda_{1,2}}\ .
1076: \ee
1077: Similarly for $\kappa_1/\kappa$,
1078: \be
1079: \n{5.12b}
1080: \frac{\kappa_1}{\kappa} \approx A_3(v)\left(\frac{1}{b}\right)^{\lambda_3}\ .
1081: \ee
1082: The values for $A_{1,2,3}$ and $\lambda_{1,2,3}$ obtained from a linear fit are shown
1083: in table~\ref{table1}.
1084: By comparing~(\ref{5.12b}) with (\ref{5.7}) one can conclude that the numerical value
1085: $\lambda_3=2.23$ is close to the value $2$, which enters the relation (\ref{5.7}).
1086: 
1087: Note that the last data point on the plot for $\kappa_2$ is a bit out of a straight line. 
1088: We are not sure if this is a genuine feature or simply inaccurate data points (e.g., due to
1089: round-off errors)\footnote{The discrepancies are larger than those which could result 
1090: from the extrapolation procedure.}.
1091: We did not include these data points into calculation of the values in table~\ref{table1}.
1092: 
1093: \begin{table}[!htb]
1094: \bea
1095: \begin{array}{|c||c|c|c|c|}\hline
1096:  v/c        & 0.462 & 0.762 & 0.964 & 0.995  \\ \hline\hline
1097: A_1         & 22.3  & 66.6  & 231.0 & 631.6  \\ \hline
1098: \lambda_1   & 2.31  & 2.36  & 2.36  & 2.35   \\ \hline
1099: A_2         & -83.9 &-381.4 &-1487.9&-3717.0 \\ \hline 
1100: \lambda_2   & 3.77  & 3.89  & 3.87  & 3.83   \\ \hline
1101: A_3         &  -    &5.42   &6.21   & 6.17 \\ \hline 
1102: \lambda_3   &  -    &2.23   &2.23   &2.23 \\ \hline
1103: \end{array}
1104: \eea
1105: \caption{Values of fitted parameters from equation~(\ref{5.12}) and (\ref{5.12b}) }
1106: \label{table1}
1107: \end{table}
1108: 
1109: 
1110: \subsection{Form of the kinks}
1111: 
1112: 
1113: \begin{figure}[!htb]
1114: \begin{tabular}{c}
1115: \epsfig{file=kink_b80_i1p0.eps, width=8cm}
1116: \hfill
1117: \epsfig{file=kink_b6_i1p0.eps, width=8cm}\\
1118: \end{tabular}
1119: \caption{Profiles of the kinks for different impact parameters}
1120: \label{kinks}
1121: \end{figure}
1122: 
1123: 
1124: Figure~\ref{kinks} shows details of the transition from the ``old phase'' 
1125: to the ``new phase'' for two different impact parameters. 
1126: We see that the width of the kinks differs significantly.
1127: In general the width also depends on the velocity $v/c$..
1128: The dependence of the kink width on velocity and the impact parameter 
1129: can be estimated from the weak field approximation as
1130: \be
1131: w \sim \frac{bc}{v}\ .
1132: \label{n5.2}
1133: \ee
1134: 
1135: We operationally define the width of the kink to be the width of the peak
1136:  of $dY/d\sigma$ at $1/20$th of its hight.
1137: Figure~\ref{kink3d} shows comparison of the kink widths obtained from 
1138: simulation with a function 
1139: 
1140: \be
1141: \n{5.13}
1142: w_{\rm analyt} = 7.14\frac{bc}{v}\ .
1143: \ee
1144: The numerical constant was obtained by a fit. The match with the data is very good.
1145: 
1146: We also noticed that the dependency of the width of the kink 
1147: on the rotation parameter $a$ is rather small.
1148: In general
1149: \be
1150: w_{a/M=1}<w_{a/M=0}<w_{a/M=-1}\ .
1151: \ee
1152: For the parameters shown on figure~\ref{kink3d} the rotation made a difference
1153:  from  $3M$ to $7M$ for the extremal black holes.
1154: 
1155: \begin{figure}[!ht]
1156: \begin{center}
1157: \epsfig{file=kink3d.eps, width=12cm}
1158: \caption{Comparison of kink widths for $v/c=0.462,0.762,0.964,0.995$ and 
1159: impact parameters $b/M=9,12,15,18,25,40$ obtained from simulation
1160: (circles) and the analytic approximation~(\ref{5.13}). The data shown are taken
1161: for $a/M=0$.}
1162: \label{kink3d}
1163: \end{center}
1164: \end{figure}
1165: 
1166: 
1167: \section{Discussion}
1168: \setcounter{equation}0
1169: 
1170: In this paper we studied the scattering of a long test cosmic string
1171: by a rotating black hole. We demonstrated that qualitatively many
1172: general features of the weak field scattering are present in
1173: scattering of the cosmic string in the strong field regime.
1174: Displacement of the string in the $Y-$direction always has the form
1175: of a transition of the string from the initial phase (initial plane) to the
1176: final phase (final plane which is parallel to the initial one and
1177: displaced in the direction to the black hole by distance
1178: $\kappa$). The boundary between these phases is a kink moving with the
1179: velocity of light away from the center. An important
1180: difference between weak- and strong-field scattering is in the
1181: dependence of $\kappa$ on the impact parameter $b$ and velocity $v$.
1182: In general $\kappa$ for strong field scattering is much
1183: greater that its value calculated by weak-field approximation.
1184: It is also always greater for retrograde scattering than
1185: for a prograde scattering. The explanation of this is quite simple: The
1186: retrograde string spends more time in the vicinity of the black hole
1187: than a prograde one. This effect is a result of the dragging of the
1188: string into rotation by the black hole.
1189: 
1190: The width of the kink for the
1191: strong field scattering is smaller than for the scattering in the weak 
1192: field regime.
1193: 
1194: As was demonstrated in previous studies for string scattering by
1195: non-rotating black holes, for any given velocity there exists a critical
1196: value $b_{crit}$ of
1197: the initial impact parameter so that for $b<b_{crit}$ the string is
1198: trapped by the black hole, while for $b>b_{crit}$ it is scattered by
1199: the black hole (and all points on the string reach infinity). In this paper we
1200: focus our attention on scattering in the regime where $b$ is
1201: significantly greater than $b_{crit}$. The study of string capture and
1202: near-critical string scattering requires modification of the
1203: computational program and more time consuming calculations. We are
1204: going to present these results in a subsequent publication.
1205: 
1206: 
1207: 
1208: \vspace{12pt} 
1209: {\bf Acknowledgments}:\ \ The authors are
1210: grateful to Don Page for various stimulating discussions.
1211: This work was partly supported by the Natural
1212: Sciences and Engineering Research Council of Canada.  One of the
1213: authors (V.F.) is grateful to the Killam Trust for its  financial
1214: support. M.S. thanks FS Chia PhD Scholarship.
1215: Our work made use of the infrastructure and resources of MACI (Multimedia Advanced
1216: Computational Infrastructure), funded in part by the CFI (Canada Foundation for Innovation),
1217: ISRIP (Alberta Innovation and Science Research investment Program), and the Universities of
1218: Alberta and Calgary.
1219: 
1220: 
1221: \appendix
1222: 
1223: \section{Numerical scheme details}
1224: \setcounter{equation}0
1225: 
1226: In section \ref{sec3.2} we briefly described the numerical scheme we adopted.
1227: In this appendix we present some more details.
1228: Figure \ref{f02} shows the overall structure of the time domain.
1229: 
1230: The numerical grid in the sigma direction is non-uniform. 
1231: This is very important since there are regions where the string is relatively straight and regions
1232: with high curvature. By far the numerically most ``sensitive'' region is the kink.
1233: Therefore we make the grid denser in the vicinity of the kink. The length of the dense zone
1234: is chosen to be approximately twice the kink width\footnote{The grid within each zone is uniform}.
1235: The steeper the kink the denser the grid must be. Typically, the dense zone is $3$--$30$ times
1236: denser then the rest of the grid.
1237: 
1238: The time step $\Delta\tau$ must not exceed the smallest distance between any two grid 
1239: points $\Delta\sigma$. A convenient choice turns out to be 
1240: \be
1241: \Delta\tau = \Delta\sigma\sqrt{1-v^2/c^2}\ .
1242: \ee
1243: This means that dense grid also implies small time step which in turn makes the 
1244: simulation time longer.
1245: The formula~(\ref{n5.2}) explains why it is more difficult to deal with small impact
1246: parameters and relativistic velocities.
1247: 
1248: Since the kink is moving away from the central point $\sigma=0$ the dense zone must move 
1249: with it. Therefore we regularly check the position of the kink and adjust the grid 
1250: when the kink reaches the boundary of the dense zone. This procedure is illustrated in figure
1251: \ref{grid}. Note that we introduced similar (but much shorter) dense zone at the edge
1252: of the string. This is to prevent from ``cutting'' the string too quickly (every time 
1253: step we lose one grid point). 
1254: When we change the the grid structure we must use interpolation to obtain the variable values 
1255: at the new grid points. In our case we used cubic spline which proved to work very well.
1256: 
1257: \begin{figure}[!h]
1258: \begin{center}
1259: \epsfig{file=fig03b.eps, width=0.75\textwidth,height=12cm}
1260: \caption{Structure of the numerical grid}
1261: \label{grid}
1262: \end{center}
1263: \end{figure}
1264: 
1265: At certain stage the two dense zones merge and eventually they disappear altogether (which 
1266: means that the kink passed the edge of our numerical domain). At this point we can increase 
1267: $\Delta\tau$ to reflect the sparser grid.
1268: 
1269: Since our problem is naturally formulated on rectangular grid the finite difference method is
1270: used to discretize the equations of motion.
1271: To approximate the first and second derivatives we use standard second order centered formulas,
1272: e.g.,
1273: \be
1274: (\pa_{\tau}{\cal X}^\mu)_{i,j} = \frac{{\cal X}^\mu_{i+1,j}-{\cal X}^\mu_{i-1,j}}{2\Delta\tau}
1275: \ee
1276: \be
1277: (\pa^2_{\tau}{\cal X}^\mu)_{i,j} = \frac{{\cal X}^\mu_{i+1,j}-2{\cal X}^\mu_{i,j}+{\cal X}
1278: ^\mu_{i-1,j}}{\Delta\tau^2}\ ,
1279: \ee
1280: where index $i$ marks the $i$-th time step and $j$ marks the $j$-th grid point in the sigma
1281: direction.
1282: Similar expressions are used for the spatial derivatives except for the points lying
1283: at the boundary between dense and ``normal'' zones. In that case we must use modified formulas
1284: \be
1285: (\pa_{\sigma}{\cal X}^\mu)_{i,j} = \frac{1}{\Delta\sigma_{j-1}+\Delta\sigma_{j+1}}
1286: \left(\frac{\Delta\sigma_{j-1}}{\Delta\sigma_{j+1}}({\cal X}^\mu_{i,j+1}-{\cal X}^\mu_{i,j})
1287: +\frac{\Delta\sigma_{j+1}}{\Delta\sigma_{j-1}}({\cal X}^\mu_{i,j}-{\cal X}^\mu_{i,j-1})\right)
1288: \ee
1289: \be
1290: (\pa^2_{\sigma}{\cal X}^\mu)_{i,j} = \frac{2}{\Delta\sigma_{j-1}+\Delta\sigma_{j+1}}
1291: \left(\frac{1}{\Delta\sigma_{j+1}}({\cal X}^\mu_{i,j+1}-{\cal X}^\mu_{i,j})
1292: -\frac{1}{\Delta\sigma_{j-1}}({\cal X}^\mu_{i,j}-{\cal X}^\mu_{i,j-1})\right)\ ,
1293: \ee
1294: where $\Delta\sigma_{j-1}=\sigma_j-\sigma_{j-1}$ and $\Delta\sigma_{j+1}=\sigma_{j+1}-\sigma_{j}$.
1295: 
1296: After the discretization process we obtain a set of four coupled non-linear equations for 
1297: the four unknown variables ${\cal X}^\mu_{i+1,j}$ at each grid point $\sigma_j$. 
1298: Since the equations for different $j$'s are independent of each other we can solve them in parallel.
1299: 
1300: 
1301: 
1302: \begin{thebibliography}{000}
1303: 
1304: \bibitem{ShVi:94}  Vilenkin, A. and Shellard, E.~P.~S., {\em Cosmic strings 
1305: and other topological defects.} (Cambridge Univ. Press, Cambridge) (1994).
1306: 
1307: \bibitem{KoLi:87} L. A. Kofman and A. D. Linde. Nucl. Phys. {\bf
1308: B282}, 555 (1987).
1309: 
1310: \bibitem{LiRi:97} A. D. Linde and A. Riotto. Phys. Rev. {\bf D56},
1311: 1841 (1997).
1312: 
1313: \bibitem{BiDePe:98} P. Bin\'{e}truy, C. Deffayet and P. Peter. Phys.
1314: Lett. {\bf B441}, 52 (1998).
1315: 
1316: \bibitem{PoVa:99} L. Pogosian and T. Vachaspati. Phys. Rev. {\bf D60},
1317: 083504 (1999).
1318: 
1319: \bibitem{Cant:00} C. R. Contaldi. E-print astro-ph/0005115 (2000).
1320: 
1321: \bibitem{SiPe:01} N. Simatos and L. Perivolaropoulos. Phys. Rev. {\bf
1322: D63}, 025018 (2001).
1323: 
1324: \bibitem{LaSh:02} M. Landriau and E. P. S. Shellard. E-print
1325: astro-ph/0208540 (2002).
1326: 
1327: \bibitem{BoPeRiSa:02} F. R. Bouchet, P. Peter, A. Riazuelo and M.
1328: Sakellariadou. Phys. Rev. {\bf D65}, 021301 (2002).
1329: 
1330: \bibitem{DVFr:97} De Villiers, J.P., and Frolov, V.P.  
1331: Int. J. Mod. Phys. {\bf D7} No. 6 (1998) 957-967.
1332: 
1333: \bibitem{DVFr:98} De Villiers, J.P., and Frolov, V.P.  
1334: Phys. Rev. {\bf D58}, 105018(8) (1998).
1335: 
1336: \bibitem{Page:98} Page, D.N. Phys. Rev. {\bf D58}, 105026(13) (1998).
1337: 
1338: \bibitem{DVFr:99} De Villiers, J.P., and Frolov, V.P.  
1339: Class. Quant. Gravity {\bf 16}, 2403   (1999).
1340: 
1341: \bibitem{Page:99} Page, D.N.  Phys.Rev. {\bf D60}, 023510 (1999).
1342: 
1343: \bibitem{Poly:81} Polyakov, A.M. Phys. Lett. {\bf B103}, 207 (1981).
1344: 
1345: \bibitem{MTW} Misner, C.W.,Thorne, K.S., and Wheeler, J.A. {\em
1346: Gravitation} (W.H.~Freeman, San Francisco, 1973), Section~40.7. 
1347: 
1348: \bibitem{ThLe:18} Thirring, H., and Lense, J. Phys. Z. {\bf 19}, 156
1349: (1918).
1350: 
1351: 
1352: \end{thebibliography}
1353: 
1354: 
1355: 
1356: \end{document}
1357: