1: \documentclass[aps,prd,nofootinbib]{revtex4}
2: %\documentclass[aps,prd,preprint]{revtex4}
3: \usepackage{graphicx}
4: \usepackage{amsmath,amssymb,latexsym}
5:
6: %\addtolength{\oddsidemargin}{-.575in}
7: %\addtolength{\evensidemargin}{-.575in}
8: %\addtolength{\textwidth}{1.15in}
9:
10: %\addtolength{\topmargin}{-0.5in}
11: %\addtolength{\textheight}{1.00in}
12:
13: \begin{document}
14: \title{Towards the Final Fate of an Unstable Black String}
15:
16: \author{Matthew Choptuik$^{1,2}$, Luis Lehner$^{3,2}$, Ignacio (I\~naki) Olabarrieta$^{2}$, \\
17: Roman Petryk$^{2}$, Frans Pretorius$^{4,2}$ and Hugo Villegas$^{2}$}
18:
19: \address{$^{1}$ CIAR Cosmology and Gravity Program \\
20: $^{2}$ Department of Physics and Astronomy\\
21: University of British Columbia,\\
22: Vancouver, CANADA V6T 1Z1 \\
23: $^{3}$ Department of Physics and Astronomy\\
24: Louisiana State University, \\
25: Baton Rouge, LA 70810.\\
26: $^{4}$ Theoretical Astrophysics 130-033\\
27: California Institute of Technology,\\
28: Pasadena, CA, 91125}
29: \begin{abstract}
30: Black strings, one class of higher dimensional analogues of black
31: holes, were shown to be unstable to long wavelength perturbations
32: by Gregory and Laflamme in 1992, via a linear analysis.
33: We revisit the problem through numerical solution of the full equations
34: of motion, and focus on trying to determine the
35: end-state of a perturbed, unstable black string. Our preliminary
36: results show that such a spacetime tends towards a solution
37: resembling a sequence of spherical black holes connected by thin
38: black strings, at least at intermediate times. However, our code
39: fails then, primarily due to large gradients that develop
40: in metric functions, as the coordinate system we use is not
41: well adapted to the nature of the unfolding solution. We are thus
42: unable to determine how close the solution we see is to the final
43: end-state, though we do observe rich dynamical behavior of the system
44: in the intermediate stages.
45: \end{abstract}
46: \maketitle
47:
48: \section{Introduction}
49: The stability of four-dimensional black holes is a well known,
50: and fundamental result of relativity theory~\cite{chandra}.
51: The picture in higher dimensional
52: spacetimes was shown to be quite different by Gregory and
53: Laflamme~\cite{gregorylaflame1,gregorylaflame2}, who demonstrated
54: the existence of unstable long wavelength modes of
55: the black string in perturbation theory.
56: This finding, coupled to arguments based on entropy considerations, led
57: to their conjecture that black strings might bifurcate into higher
58: dimensional analogues of spherical black holes.
59: Cosmic censorship would be violated were this the case,
60: since a singularity must be
61: encountered by a bifurcating black hole horizon, essentially as a
62: consequence of the principle of equivalence~\cite{hawkingellis}.
63:
64: The existence of the Gregory-Laflamme instability has been assumed in
65: many subsequent studies of higher dimensional gravity theory, including
66: the classical limit of string theory
67: (see, for example, \cite{hms96,imsy98,agmoo00,hv00}), some of which
68: have also assumed the validity of the bifurcation conjecture.
69: However, a linearized analysis can say little, if anything, concerning
70: the nature of the full non-linear evolution of an unstable string, and
71: the final end-state of such a configuration remained to be established.
72:
73: Recently, Horowitz and Maeda were able to prove, under some
74: assumptions, that black strings cannot bifurcate in finite
75: time~\cite{horowitzmaeda}. Furthermore, they conjectured that the
76: system is likely to approach a new stationary solution which is
77: not translationally invariant along the string direction. However, even if
78: the assumptions involved in the proof are sufficiently generic, their
79: analysis cannot identify the final end-state of evolution.
80: Partial answers can be
81: sought via perturbation analysis as done by Gubser~\cite{gubser}.
82: By assuming the Horowitz-Maeda conjecture and linearizing
83: the solution {\em at the critical length} to first order (and a
84: partial extension to second order), Gubser argued that the transition
85: to the final solution must be of second order type (i.e.
86: discontinuous). Despite these developments, it seems clear that a convincing
87: answer to the question at hand
88: can only be obtained by solving the full equations governing the
89: problem. A step in this direction would be to search for special
90: solutions, such as stationary ones, and compare the physical content of the
91: obtained configurations with the black string solutions. This has recently
92: been carried out by Wiseman~\cite{wiseman}, who numerically
93: solves the equations resulting from a static ansatz.
94: Interestingly, he finds non-uniform solutions with mass larger than
95: that of the black string for a given compactification radius.
96: Thus, he concludes that the solutions he finds
97: cannot be the end-states
98: conjectured by Horowitz and Maeda (also see related work by Kol~\cite{kol}).
99:
100: Additional work by Unruh and Wald~\cite{unruhwald} studies the dynamics of a
101: uniform cylindrical configuration of matter in Newtonian gravity. They observe
102: that a perturbation of the density gives raise to a Jeans instability
103: responsible for the collapse of the system along the cylinder's length.
104: They then argue that if the main features of this model are robust in the passage to the
105: general relativistic system, one possible end-state for the perturbed black
106: string would be collapse in the string direction, resulting in singularity
107: formation.
108: Note that this collapse need not lead to violations
109: of cosmic censorship, as the final
110: singularity could still be hidden by an event horizon~\cite{geddes}.
111:
112: Clearly there are several distinct viable possibilities for the final
113: end-state of a perturbed black string, with
114: remarkably different
115: consequences associated with the range of options.
116: Current conjectures range from ``nothing interesting happens'', to
117: violations of cosmic censorship, to the arguably more extreme case of
118: a complete collapse of the
119: spacetime. In order to completely settle the issue,
120: the full dynamics of the perturbed black string needs to be addressed.
121: At least in principle, this will allow us to identify which of the
122: above possibilities (if any!) is actually realized.
123: In this paper we report on preliminary work in this
124: direction---a program to simulate the dynamics of the black string
125: through numerical solution of the Einstein equations.
126: At this stage of the
127: project, we cannot yet provide an answer to the question of the
128: end-state; however, we have tantalizing
129: results that show the spacetime going through a configuration
130: resembling 5-dimensional spherical black holes connected by thin
131: black strings that expand along the string dimension. Our simulations
132: eventually crash while the spacetime is still fairly dynamical,
133: and so we cannot determine whether what we see is near the end-state, or
134: merely an intermediate configuration in a more complicated
135: evolution. Underlying the current failure of our simulations is the
136: fact that the coordinate system employed is not well adapted to the solution
137: that unfolds at late times, wherein fatally steep gradients develop in
138: metric functions~\footnote{This divergence of metric gradients
139: does not happen earlier as
140: resolution is increased, and is not accompanied by divergence of curvature
141: invariants such as the Kretschmann scalar. This suggests that the code is
142: evolving to a coordinate, rather than a geometric, singularity.}.
143:
144: The outline of the remainder of the paper is as follows.
145: In Sec.~\ref{sec:setup},
146: we begin by describing the equations of motion, our
147: coordinate choices, generation of initial data, as well as our numerical solution scheme.
148: Additionally, we also briefly mention the tools
149: we employ to monitor the solution, deferring details to the Appendices.
150: In Sec. \ref{sec:results} we discuss the results obtained with this code, and
151: conclude in Sec. \ref{sec:conclusion} by mentioning directions for future work
152: that may allow us to more definitively answer questions regarding the
153: end-state of the Gregory-Laflamme instability.
154:
155: \section{Equations, Black Strings and Numerics}\label{sec:setup}
156: We wish to solve the vacuum Einstein equations in higher dimensional settings.
157: For simplicity, and without loss of
158: generality in studying the Gregory-Laflamme instability, we only consider
159: the 5-dimensional case, and restrict attention
160: to spherical symmetry within the 4-dimensional subspace tangent
161: to the ``extra'' dimension. We also use the natural generalization of the
162: ADM decomposition to derive the system of equations that we
163: then solve numerically.
164: Choosing units in which $G=c=1$, and adopting MTW~\cite{MTW} conventions,
165: our starting point is thus a metric element given by
166: \begin{eqnarray}\label{metric}
167: ds^2=(-\alpha^2+\gamma_{AB} \beta^A \beta^B) dt^2 + 2 \gamma_{AB} \beta^A dx^B
168: dt + \gamma_{AB} dx^A dx^B + \gamma_{\Omega} d\Omega^2
169: \end{eqnarray}
170: where $x^A=(r,z)$, and $d\Omega^2$ is the 2-spherical line element
171: with coordinates chosen orthogonal to the $t=$ constant
172: congruences (hence there is no shift corresponding to angular
173: directions). All metric components defined via (\ref{metric}) depend upon
174: $(t,r,z)$: $t$ is a time-like coordinate, $r$ is a radial coordinate,
175: and $z$ is the coordinate along the length of the string. To further
176: expedite the numerical implementation,
177: we make $z$ a periodic coordinate
178: by identifying $z=0$ and $z=L$. Then, following the results
179: of Gregory and Laflamme, we expect black strings
180: to be unstable only for $L$ greater than some critical length $L_c$.
181:
182: The vacuum Einstein equations, written in ADM form \cite{MTW,york_78},
183: are 1) the Hamiltonian constraint
184: \begin{equation}\label{HC}
185: H \equiv ^{(4)}R + K^2 - K_{ab} K^{ab} = 0,
186: \end{equation}
187: 2) the momentum constraints
188: \begin{equation}\label{MC}
189: M_a \equiv K_{a\ \ |b}^{\ b} - K_{|a} = 0,
190: \end{equation}
191: 3) the evolution equations for the $\gamma_{ab}$
192: \begin{equation}\label{gdot}
193: \frac{\partial \gamma_{ab}}{\partial t} = - 2 \alpha K_{ab}
194: + \beta_{a|b} + \beta_{b|a},
195: \end{equation}
196: that follow from the definition of the extrinsic curvature
197: $K_{ab}$ associated with $t={\rm const.}$ slices,
198: and 4) the evolution equations for the extrinsic curvature
199: \begin{eqnarray}\label{kdot}
200: \frac{\partial K_{ab}}{\partial t} &=& \alpha \left(^{(4)}R_{ab} +
201: K K_{ab}\right) -2\alpha K_{a c} K^c{}_{b} -\alpha_{|ab} \nonumber
202: \\& & + \beta^c{}_{|a} K_{cb} + \beta^c{}_{|b} K_{ca} +\beta^c K_{ab|c}
203: + \alpha F^c_a F^d_b \gamma_{cd} H .
204: \end{eqnarray}
205: In the above, $a,b,\ldots$ are four-dimensional (spatial) indices,
206: $^{(4)}R_{ab}$ and $^{(4)}R$ are, respectively, the
207: Ricci tensor and Ricci scalar intrinsic to the four-dimensional spatial
208: hypersurfaces, $\alpha$ is the lapse function, $\beta^c$ is
209: the shift vector, the vertical bar $_|$ denotes covariant
210: differentiation in the spatial hypersurfaces (compatible with $\gamma_{ab}$),
211: and $F^c_a = -2 \delta^c_r \delta^r_a$. We note that the term proportional
212: to $H$ in~(\ref{kdot}) has been added as a result of stability considerations;
213: see for instance, the discussions in~\cite{lehnerreview,shinkai,kelly}.
214: To simplify the final set of
215: equations solved numerically, as well as to regularize certain terms that
216: otherwise diverge
217: at spatial infinity (see the discussion of our coordinate system
218: in the next sub-section), we define the following variables
219: \[
220: g_{rr} \equiv \gamma_{rr} \quad\quad\quad g_{rz} \equiv \gamma_{rz}
221: \quad\quad\quad g_{zz} \equiv \gamma_{zz}
222: \]
223: \[
224: g_{\theta\theta} \equiv \gamma_{\Omega}/r^2
225: \quad\quad\quad g_{\phi\phi} \equiv \gamma_{\Omega}/(r^2\sin^2\theta)
226: \]
227: \[
228: k_{rr} \equiv r^2 K_{rr}/\alpha \quad\quad\quad k_{rz} \equiv K_{rz}
229: \quad\quad\quad k_{zz} \equiv K_{zz}
230: \]
231: \begin{equation}
232: k_{\theta\theta} \equiv K_{\theta\theta}/\alpha
233: \quad\quad\quad k_{\phi\phi} \equiv K_{\theta\theta} / (\alpha \sin^2\theta)
234: \end{equation}
235: and use them as the fundamental dynamical quantities in our numerical code.
236: As discussed in the following sub-sections,
237: to complete the prescription of the evolution problem
238: we need to choose a suitable lapse and shift, specify
239: initial and boundary conditions, and then implement these choices
240: and specifications consistently.
241:
242: \subsection{Boundary and Coordinate Conditions}
243: A particular concern here is that ``standard'' outer boundary
244: conditions~\cite{lehnerreview}, often imposed during numerical
245: evolution of Einstein's equations,
246: might not be well suited for studying the string
247: instability. In particular, we must
248: be able to evolve for very long times while absolutely minimizing spurious
249: influences from the outer boundary of the computational domain.
250: In addition, in the present case
251: we cannot assume, {\em a priori}, that any given initial configuration will
252: settle down to some stationary solution; thus, boundary conditions predicated
253: on such assumptions (such as a $1/r$ fall-off condition in a metric
254: component), when imposed at a finite proper
255: distance from the black string, could very well adversely affect the
256: numerical results~\footnote{In fact, in an earlier version of the code that did
257: not use a radially compactified coordinate system, we did
258: encounter such problems, in that some artificial stationary
259: non-homogeneous solution
260: was apparently entirely ``sourced'' via an outer boundary located at
261: a finite distance from the string.}. To
262: ensure minimal boundary influence we therefore extend the domain
263: of integration to $i^o$ by radially compactifying the spacelike
264: hypersurfaces via the introduction of a new coordinate, $x$, defined by
265: \begin{equation}
266: x\equiv \frac{r}{1+r} \, .
267: \label{defx}
268: \end{equation}
269: As might be expected, this transformation causes computational problems of
270: its own---most notably decreased spatial resolution at large distances---but,
271: as discussed in Sec.~\ref{numerics}, we can deal with these
272: difficulties using numerical dissipation.
273: Having introduced the new compactifying coordinate, we can directly impose
274: boundary conditions derived from the demand of asymptotic flatness at
275: spatial infinity, which lies at $x=1$.
276:
277: We employ singularity excision techniques~\cite{unruh} to allow us
278: to evolve the entire perturbed black string spacetime exterior to
279: the apparent horizon (plus a certain ``buffer zone'' that lies within
280: the horizon).
281: Hence, we do not need to impose inner boundary conditions as long as the
282: $t=\rm{const.}$ hypersurfaces penetrate the horizon, and that all
283: characteristics of the evolution equations are in-going on the
284: boundary.
285: Ensuring that this is the case involves choosing ``good''
286: coordinate conditions (choice of lapse and shift), which, for generic
287: string evolutions, remains an open problem.
288: As a preliminary step, we have based our coordinate choices on
289: those that yield the ingoing Eddington-Finkelstein form
290: of the unperturbed black string metric
291: \begin{equation}\label{bs_metric}
292: ds^2_{\rm BS}= -(1-2M/r) dt^2 + 4 M/r dr dt + (1+2M/r) dr^2 + dz^2 +
293: r^2 d\Omega^2 \, .
294: \end{equation}
295: Comparison with the general 5-dimensional ADM form provides the
296: identifications
297: \begin{eqnarray}
298: \alpha_{\rm BS} &=& (1+2M/r)^{(-1/2)}\, , \label{bs_lapse}\\
299: \beta_{\rm BS}^A &=& (2M/(r+2M))\delta^A_r\,. \label{bs_shift}
300: \end{eqnarray}
301: For reference we
302: also list the two non-trivial components of the
303: extrinsic curvature of a $t={\rm const.}$ slice defined by (\ref{bs_metric}):
304: \begin{eqnarray}\label{bs_K}
305: K_{r r} &=& -2M \frac{(r+M)}{r^3} \sqrt{\frac{r}{r+2M}}, \nonumber \\
306: K_{\theta \theta} &=& 2M \sqrt{\frac{r}{r+2M}}.
307: \end{eqnarray}
308: In generalizing~(\ref{bs_lapse}) and (\ref{bs_shift}) to the dynamical case,
309: we have chosen $\alpha=\alpha_{\rm BS}$ and $\beta^{z}=\beta_{\rm BS}^{z}=0$. In a
310: preliminary version of our code, we also required that
311: $\beta^{r}=\beta_{\rm BS}^{r}$. This, however, caused a coordinate
312: pathology to develop at late times during the evolution of
313: unstable strings---specifically, some regions of the horizons
314: approached a zero coordinate-radius, while maintaining {\em finite}
315: proper radius.
316: In our current efforts, we
317: choose $\beta^{r}$ such that $g_{\theta \theta}$ remains constant
318: during evolution \cite{seidelexcision,lehnercadm}, by requiring that
319: \begin{equation}\label{br_gauge}
320: \beta^r = \frac{2 \alpha K_{\theta \theta}}{{\gamma_{\theta \theta}}_{,r}}.
321: \end{equation}
322: This shift condition performs reasonably well, as will
323: be seen in Sec. \ref{sec:results}. However, our current simulations still
324: suffer from ``grid-stretching'' problems in the $z$ direction at late times,
325: suggesting that a more dynamical gauge condition for $\beta^z$
326: (and possibly for $\alpha$ and $\beta^{r}$) could be useful. This issue
327: is discussed in more detail in Sec.~\ref{sec:conclusion}.
328:
329: \subsection{Initial Data \label{ID}}
330: As anticipated,
331: we observe that even numerical truncation errors,
332: if non-uniform in the $z$ direction, are enough to
333: trigger the Gregory-Laflamme instability in our simulations.
334: However, to reduce the computational effort required to reach the
335: ``interesting'' (i.e. non-perturbative) stages of
336: evolution, we adopt initial configurations whose departure from
337: the black string solution can be arbitrarily tuned. In order to
338: find such data, we must solve the Hamiltonian constraint, and the $r$
339: and $z$ components of the momentum constraint (the other
340: components of the momentum constraint are trivially satisfied
341: because our coordinate system is adapted to spherical symmetry).
342: The deviation---not necessarily
343: small---from the black string solution, is introduced
344: via $g_{\theta \theta}$, and takes the following form:
345: \begin{equation}
346: \label{gtt}
347: g_{\theta \theta}(0,r,z) = 1 + A\sin{\left(z \frac{2 \pi q}{L}\right)}
348: e^{ -(r-r_o)^2 / \delta_r^2 }.
349: \end{equation}
350: Here, $A$ is used to control the overall strength of the ``perturbation'',
351: while $q$ is an
352: integer that controls the spatial frequency in the $z$ direction.
353: For the results presented below, $A=0.1$, $q=1$, $r_0=2.5$ and
354: $\delta_r=0.5$, and we perturb about a unit mass ($M=1$)
355: black string solution.
356: As described in more detail in Appendix \ref{app:initialdata},
357: $g_{rr}$, $k_{rr}$, $k_{\theta \theta}$ are then calculated by
358: solving the constraint equations, with the remainder
359: of the metric and extrinsic curvature variables set
360: to the values they would take for an unperturbed black string
361: (see (\ref{bs_metric}) and (\ref{bs_K})).
362:
363: \subsection{Numerical Evolution\label{numerics}}
364:
365: To numerically evolve the initial data sets described above,
366: we discretize the evolution equations~(\ref{gdot})-(\ref{kdot}) using
367: second-order accurate finite difference techniques that include
368: Crank-Nicholson treatment of the temporal and spatial derivatives.
369: We use a uniform distribution of grid points in $z$ and $x$
370: (recall that $x\equiv r/(1+r)$).
371: The resulting implicit system of algebraic equations is solved
372: iteratively.
373: We initially implemented a serial version of the algorithm, and later
374: coded a parallel version using the CACTUS Computational Toolkit~\cite{cactus},
375: wherein the equations of motion, monitoring tools and I/O were handled by our
376: own routines, suitably interfaced to CACTUS.
377:
378: Black hole excision is handled as follows. We periodically find
379: the apparent horizon, as discussed in the following
380: sub-section and Appendix \ref{app:AH}. We then define the surface
381: along which we excise to be a certain number of ``buffer'' points
382: {\em inside} the apparent horizon
383: (typically $10$---$30$ buffer points are used)~\footnote{In several
384: tests, we also adopted an excision region given by the global minimum
385: $r$ value of the apparent horizon and compared the results with those obtained
386: when the excision region was defined by the apparent horizon.
387: The agreement obtained
388: gives extra indication that the excision implementation is consistent.}.
389: During each Crank-Nicholson iteration, all the evolution
390: difference equations are applied up to the excision surface, and
391: any function values referenced by finite difference stencils
392: interior to this surface are defined via fourth order
393: extrapolation. When the apparent horizon location, and hence excision surface
394: changes during evolution, function values at all repopulated
395: points (i.e. those that moved from inside to outside the excision
396: surface during the time step) are computed via the same fourth
397: order extrapolation routine. The one exception to this procedure is for
398: the grid values of
399: $g_{\theta\theta}$, which we specify {\em a priori} on the
400: entire computational domain, and that remain fixed due to our gauge
401: choice (\ref{br_gauge}). Moreover, we have found it useful to
402: choose a functional form for $g_{\theta\theta}$ that tends to zero
403: at some positive value of $r$ (though inside the original apparent horizon
404: location and outside of the limits of integration of the initial
405: data).
406: This causes the ``pinching-off'' of the unstable black
407: string to be less severe in coordinate space, i.e. we approach
408: zero areal radius at a finite coordinate $r$.
409: In turn, this slightly reduces the virulence of the coordinate problems we
410: observe at late times, and also provides better
411: load-balancing of the parallel code, given the method CACTUS uses to
412: distribute grids among processors.
413:
414: For the evolution, we choose a time step $dt=\lambda_{{\rm CFL}} {\rm min}(dr,dz)$, where
415: the constant $\lambda_{{\rm CFL}}$ must be set less than $1/\sqrt{2}$ in order
416: to satisfy the Courant-Friedrichs-Lewy (CFL) stability condition that
417: results from our iterative solution of the Crank-Nicholson scheme (typically
418: we use $\lambda_{{\rm CFL}}=0.25$). Note that this restriction on $\lambda_{{\rm CFL}}$ is
419: based upon flat-space light speeds within our coordinate system ((\ref{bs_metric})
420: with $M=0$), which, for the solutions presented here, are always greater than
421: or equal to the actual coordinate light speeds.
422: The function ${\rm min}(dr,dz)$
423: is calculated by only considering mesh spacings within
424: the non-excised portion of the coordinate domain. Thus, as the excision
425: surface moves, $dt$ changes with time since our grid is uniform in $x$, and
426: hence non-uniform in $r$.
427:
428: Crucially,
429: we add Kreiss-Oliger-style \cite{kreissoliger} numerical
430: dissipation to the evolution equations to control unphysical
431: high-frequency solution components (``noise'') that would otherwise
432: arise during the simulations. This is particularly helpful at the
433: excision surface, and near $i^0$, where the radial
434: compactification of points causes all outgoing wave-like
435: components of variables to eventually become poorly
436: resolved. Smoothing of the high frequency components via the
437: Kreiss-Oliger dissipation---which only targets wavelengths of size on
438: the order of the mesh spacing---prevents them from inducing numerical
439: instabilities near the outer boundary.
440:
441: \subsubsection{Monitoring the Evolution}\label{sec_monitor}
442:
443: To elucidate the nature of our computed spacetimes, we monitor the
444: following
445: quantities: 1) the location of the apparent horizon (which is also used
446: for excision as explained earlier), 2) the trajectories of null geodesics
447: that, to a certain extent, should trace the event horizon, and
448: 3) the Kretschmann invariant $I$ (the square of the Riemann tensor):
449: \begin{eqnarray}\label{I}
450: I = R_{\alpha\beta\gamma\delta}R^{\alpha\beta\gamma\delta}.
451: \end{eqnarray}
452:
453: If cosmic censorship holds---and results from our current simulations
454: provide no evidence to the contrary---then any apparent horizon found
455: will always be inside an event horizon. As is well known, although the apparent horizon can
456: often be used as a reasonable approximation to the event horizon, the
457: two do {\em not}, in general, coincide~\footnote{Indeed, depending
458: on the slicing an apparent horizon need not exist at
459: all~\cite{waldah}.}. Clearly, the event horizon is the quantity of interest in studying
460: the Gregory-Laflamme instability, and therefore we would like to
461: locate it, or at least a reasonably good approximation to it, in
462: our simulation results. Such an approximation
463: can be obtained by looking for the boundary of the causal
464: past of some $r={\rm const.}$ surface that is sufficiently far
465: outside the apparent horizon that it is certain not to be inside the event horizon, yet
466: close enough to the apparent horizon that its causal past, tracing backwards
467: from the end-time of the simulation, probes the region of interest
468: of the spacetime. We use a method to find the
469: approximate event horizon discussed by Libson et al. \cite{eh}. The approach is
470: based on radial outgoing null geodesics; as explained in \cite{eh}, the stable direction
471: for the integration of null rays that emanate
472: from the vicinity of an event horizon is backwards in time $t$. Thus, once
473: we have the complete data from the entire evolution, we start with data
474: from the latest time step available, and trace the null rays
475: backwards in time.
476:
477: Monitoring curvature scalars is useful in obtaining coordinate
478: independent information about a numerical solution of the Einstein
479: equations. In particular, $I$, as defined by~(\ref{I})
480: evaluates to $I_{\rm BS}=48M^2/\gamma_{\theta\theta}{}^3$
481: for the unperturbed black string solution, and as
482: $\gamma_{\theta\theta}$ is an invariant in spherical symmetry
483: (i.e. it is proportional to the area of an $r={\rm const.}$
484: 2-sphere), we can compare $I_{\rm BS}$ to the values of $I$ computed from
485: a numerical solution to get some indication of how close the computed solution
486: is to a black string spacetime. Furthermore, we can examine $I$ to
487: see whether curvature singularities (other than the central $r=0$
488: singularity) may be forming prior to the demise of the simulation that
489: invariably occurs when sufficiently steep metric
490: gradients develop. We note, however, that if $I$ does {\em not} diverge, it
491: does not necessarily follow that the geometry is remaining non-singular;
492: we would need to examine a larger set of curvature
493: scalars to be certain that the solutions are remaining free of
494: physical singularities.
495:
496: Appendix \ref{app:AH} contains an explanation of the method we used to find
497: apparent horizons, while
498: details of the integration techniques aimed at approximately locating event
499: horizons can be found in Appendix \ref{app:null}.
500:
501: \section{Results}\label{sec:results}
502: In this section we present results from our preliminary study of the
503: black string
504: instability. After briefly showing that we recover some of the key
505: Gregory-Laflamme results in
506: the next sub-section, we present a detailed analysis of a typical unstable
507: case in Sec. \ref{sec:unstable}. In the following, we will use the value of
508: $L_c\approx14.3M$ (with $M=1$) found by Gubser \cite{gubser},
509: which is more accurate than the value we can estimate from the
510: zero-crossing of the (positive mode) interpolating curve presented
511: in~\cite{gregorylaflame1}.
512:
513: \subsection{Recovery of Gregory-Laflamme Results}
514: We ran a variety of simulations of black strings that were perturbed
515: according to the prescription discussed in
516: section \ref{ID}. We concentrated on cases with $L$ ranging from $0.6 L_c$ to $1.8 L_c$
517: and defined $g_{\theta\theta}$ via equation (\ref{gtt}).
518: In general, we observed the expected instability for $L>L_c$, though
519: for the maximum resolution at which we performed this survey
520: ($800$ grid points in $r$
521: and $200$ points in $z$), we could only confirm $L_c$ to within about
522: $2\%$ of the expected value. In this regard we note that as
523: $L$ approaches $L_c$ from
524: above, the growth rate of the instability goes to $0$, requiring
525: longer time integrations to identify the instability, which, in turn,
526: demands ever increasing resolution
527: to counter the effects of accumulating numerical errors. Furthermore,
528: the initial configurations we have adopted contain energy in the
529: form of gravitational waves,
530: and some of this energy falls into the string early on during the evolution.
531: The increase in the mass of the string (based upon the increase in area of the apparent
532: horizon) is typically around $0.3-0.5\%$, and we would have to take this into
533: account were we to attempt to determine $L_c$ from our simulations
534: to a higher degree of accuracy.
535:
536: As a demonstration of the ability of our code to ``bracket'' the instability,
537: and following the notation of~\cite{gubser},
538: Fig. \ref{lambda_comp} shows a plot of, $\lambda$, defined by
539: \begin{equation}\label{lambda}
540: \lambda=\frac{1}{2}\left(\frac{R_{{\rm max}}}{R_{{\rm min}}}-1\right)
541: \end{equation}
542: for $L=0.975L_c$ and $L=1.03L_c$.
543: In the above, $R_{{\rm max}}$ and $R_{{\rm min}}$ are the maximum and minimum
544: areal radii, respectively, of the
545: apparent horizon at some $t={\rm const.}$ slice of the spacetime.
546: In particular, we have $\lambda=0$ for
547: the static black string spacetime.
548:
549: \begin{figure}
550: \begin{center}
551: \includegraphics[width=15cm,clip=true]{max_min_R_AH_comp.eps}
552: \end{center}
553: \caption{The maximum ($R_{{\rm max}}$) and minimum ($R_{{\rm min}}$) areal radii,
554: and the corresponding function $\lambda$ of the
555: apparent horizon as a function of time, from the evolution of
556: perturbed black strings with
557: $L=1.03L_c$ and $L=0.975L_c$. The initial
558: fluctuation in the plots correspond to the effect of the initial
559: gravitational wave perturbation, most of which either falls into
560: the string, or escapes to infinity. This close to the threshold $L_c$,
561: the growth/decay of the remnant perturbation is quite slow, and so
562: we cannot feasibly (at the resolution of the these simulations---$800 \times 200$ points in $r \times z$) follow the evolution for much
563: further than shown while maintaining reasonable accuracy (though we see no
564: signs of numerical instabilities in the stable case, and such simulations have
565: been followed to $10,000M$).
566: However, the main purpose
567: of this figure is to demonstrate the qualitative recovery of
568: the expected threshold behavior for the onset of the instability at
569: $L=L_c$.
570: \label{lambda_comp}}
571: \end{figure}
572:
573: \subsection{Beyond the Linear Regime}\label{sec:unstable}
574: We now present more detailed results from the simulation of an unstable
575: black string evolution. Specifically, we take
576: $L=1.4L_c$, since it is expected that this particular range for the $z$
577: coordinate will yield something close to the fastest growth rate for the
578: shortest wavelength instability \cite{gregorylaflame1}.
579: Because we are now probing uncharted territory with our computations,
580: we rely on convergence tests to provide an intrinsic measure of the
581: level of error in our calculations.
582: To that end,
583: we ran the simulation at several resolutions ($n_r \times n_z$): $200 \times 50$
584: (grid spacing $h$), $400 \times 100$ ($h/2$), $800 \times 200$ ($h/4$),
585: and $1600 \times 400$ ($h/8$).
586: Due to our use of a compactified radial coordinate,
587: the lowest resolution calculation cannot adequately resolve the
588: late-time behavior of the solution. However, for the ``medium resolution''
589: computation with mesh spacing $h/4$, we are apparently within the
590: convergent regime---see Fig. \ref{lambda_1000}
591: below for plots of the maximum and minimum areal radii of the apparent horizon as a
592: function of time, as well as the quantity $\lambda$ defined by
593: (\ref{lambda}), and Fig. \ref{hc_1000} for plots of the norm
594: of the Hamiltonian constraint as a function of resolution. Therefore,
595: unless otherwise noted, all the results shown below are taken from the $h/4$
596: simulation.
597:
598: \begin{figure}
599: \begin{center}
600: %%BoundingBox: 18 164 592 698
601: \includegraphics[width=15cm,clip=true]{max_min_R_AH_1000.eps}
602: \end{center}
603: \caption{The maximum ($R_{{\rm max}}$) and minimum ($R_{{\rm min}}$) areal radii,
604: and the corresponding function $\lambda = (R_{\rm max}/R_{\rm min} - 1)/2$,
605: of the apparent horizon, as a function of time, from the evolution of a
606: perturbed black string with $L=1.4L_c$.
607: $h$ labels grid spacing; hence smaller $h$ corresponds to higher
608: resolution. This plot, combined with the results shown in Fig.~\ref{hc_1000}
609: suggest that the code is in the convergent regime---in particular at later
610: times---for the $h/2$ and higher resolution
611: simulations.
612: \label{lambda_1000}}
613: \end{figure}
614:
615: \begin{figure}
616: \begin{center}
617: \includegraphics[width=14cm,clip=true]{l2norm_HC_bs_1000.eps}
618: \end{center}
619: \caption{The logarithm of the $\ell_2$-norm of the Hamiltonian
620: constraint as a function of time, evaluated on the portion of the computational domain
621: lying exterior to the apparent horizon, and from simulations
622: at several resolutions of a perturbed black string with $L=1.4L_c$.
623: As with Fig.~\ref{lambda_1000}, this plot provides evidence that
624: convergence is quite good for the $h/2$ and higher resolution
625: simulations
626: (at least until very close to when the supposed coordinate
627: singularity forms, near $t=165$).
628: \label{hc_1000}}
629: \end{figure}
630:
631: Fig. \ref{ah_embed_1000} shows embedding diagrams of the apparent
632: horizon at several times during evolution of the string,
633: and Fig. \ref{ah_embed_1000_length} shows the proper length of one period
634: of the apparent horizon (suppressing the angular coordinates) versus time.
635: (Our embedding uses the vertical axis to represent the areal radius of the apparent
636: horizon---the horizontal axis is then uniquely determined by requiring that
637: the length of the curve be equal to the proper length of the horizon).
638: The simulation crashes shortly after the last time frame shown,
639: apparently due to the coordinate pathologies that have been
640: discussed previously.
641: The
642: embedding diagrams suggest that, at least in the
643: vicinity of the apparent horizon, the solution is tending towards a
644: spacetime that can be
645: described as a sequence of spherical black holes connected by thin
646: black strings.
647: Additional, quantitative, evidence for this conjecture can be obtained
648: through a computation of the curvature invariant, $I$, on the apparent
649: horizon. For an exact black string solution, this quantity, which
650: we denote $I^0_{\rm BS}$,
651: is
652: \begin{equation}
653: I^0_{\rm BS} = \frac{12}{R_{\rm AH}^4},
654: \end{equation}
655: while for the 5-dimensional spherical black hole, the equivalent quantity,
656: $I^0_{\rm BH}$, is
657: \begin{equation}
658: I^0_{\rm BH} = \frac{72}{R_{\rm AH}^4},
659: \end{equation}
660: where, in both of the above expressions,
661: $R_{\rm AH}$ is the areal radius of the horizon. In
662: Fig. \ref{I_AH_1000} we plot the normalized quantity
663: \begin{equation}
664: I^{0N}\equiv \frac{I}{I^0_{\rm BS}} = \frac{I R_{\rm AH}^4}{12}
665: \end{equation}
666: evaluated on the apparent horizon of our
667: numerical solution of the unstable spacetime---$I^{0N}$ is $1$
668: for a black string, and $6$ for a black hole. The figure shows that,
669: as judged by $I^{0N}$,
670: the part of the apparent horizon that is forming a long neck always resembles
671: a black string---the part that is forming
672: a bulge, however, has a value of $I^{0N}$ tending towards that
673: corresponding to a black hole.
674: At $R_{{\rm max}}$, $I^{0N}$ has only reached $\sim 5$ by the time the
675: simulation ends; however, the behavior
676: of $R_{{\rm max}}$ seen in Fig.
677: \ref{lambda_1000} suggests that the growth in the normalized curvature
678: invariant, though slowing down, should continue.
679: Fig. \ref{I_AH_1000} also demonstrates the grid-stretching problems
680: that we surmise are causing the code to eventually crash---in that plot we
681: use the coordinate $z$ as the horizontal axis, and observe that the
682: relatively small region where $I^{0N}\approx1$ corresponds to the long
683: neck in Fig. \ref{ah_embed_1000}.
684: In particular, in the vicinity of the ``neck'', $g_{zz}$ becomes
685: quite large, as do its derivatives.
686:
687: Finally, in Fig. \ref{ah_eh} we show plots of the approximate event
688: horizon (as described in Sec. \ref{sec_monitor}), together with the apparent horizon
689: for the simulation. The results shown in the plot suggest that our computed
690: apparent horizon is an excellent approximation to the event horizon, at
691: least at early times
692: (not much can be said regarding the late time behavior of the event horizon, as the spacetime
693: has not settled down to a stationary state when the simulation ends).
694:
695: \begin{figure}
696: \begin{center}
697: %%BoundingBox: 18 144 502 718
698: \includegraphics[width=15cm,clip=true]{AH_embed_L2.eps}
699: \end{center}
700: \caption{Embedding diagrams of the apparent horizon, with the two
701: angular dimensions $\theta$ and $\phi$ suppressed,
702: from the $h/4$ evolution of a perturbed black string with $L = 1.4 L_c$.
703: These plots thus describe the intrinsic geometry of the apparent horizon,
704: at the given instants of constant $t$, in a coordinate system with
705: metric $ds^2=d\bar{r}^2 + d\bar{z}^2$. Here, $\bar{z}$ is a periodic coordinate,
706: and $\bar{r}$ is the areal radius of $\bar{z}=constant$ sections of the horizon.
707: To better illustrate the dynamics of the horizon, we have extended the
708: solution using the $\bar z$-periodicity, showing roughly two periods of the solution.
709: See Fig. \ref{ah_embed_1000_length} for a plot of the length of one period
710: of the apparent horizon versus time.
711: \label{ah_embed_1000}}
712: \end{figure}
713:
714: \begin{figure}
715: \begin{center}
716: %%BoundingBox: 18 164 592 698
717: \includegraphics[width=15cm,clip=true]{embed_length.eps}
718: \end{center}
719: \caption{The proper length of the apparent horizon curve in the $(r,z)$ plane
720: (between $z=0$ and $z=L$)
721: as a function of time, from the $h/4$ evolution of a perturbed black string
722: with $L = 1.4 L_c$.
723: \label{ah_embed_1000_length}}
724: \end{figure}
725:
726: \begin{figure}
727: \begin{center}
728: %%BoundingBox: 18 144 502 718
729: \includegraphics[width=15cm,clip=true]{I0_1000.eps}
730: \end{center}
731: \caption{The normalized Kretschmann invariant $I^{0N}\equiv I R_{\rm AH}^4/12$ (\ref{I}),
732: evaluated on the apparent horizon of the
733: perturbed black
734: string spacetime with $L=1.4L_c$ ($h/4$),
735: at the same times as shown in the embedding diagram
736: plots (Fig. \ref{ah_embed_1000}). Note however, that here the horizontal axis
737: is the {\em coordinate} $z$, and in particular the flat region of the
738: curve between $z\approx3.5$ and $z\approx 6.5$ in the last frame
739: corresponds to the long, thin neck region shown in the embedding diagram plot.
740: This demonstrates the rather severe ``grid-stretching'' problems we have then.
741: For the static black string spacetime, $I^{0N}=1$
742: (shown for reference as a dotted line in the figure),
743: while for a static 5D spherical black hole it evaluates to 6. This diagram
744: therefore also supports the conclusion that at late (simulation) times the
745: solution is tending
746: towards a configuration describable as a sequence of black holes connected
747: by thin black strings.
748: \label{I_AH_1000}}
749: \end{figure}
750:
751: \begin{figure}
752: \begin{center}
753: \includegraphics[width=20cm,clip=true]{ah_eh.ps}
754: \end{center}
755: \caption{Plots of the apparent horizon (labeled AH) and estimates
756: of the event horizon location (C1, C2 and C3)
757: in coordinate space (in contrast to the embedding coordinates used in
758: Fig. \ref{ah_embed_1000}), from the evolution of a
759: perturbed black string with $L=1.4L_c$, computed at resolution
760: $h/4$. Here, the C1 (C2) curve marks the inward-directed past light-cone
761: of the surface $r=10$ ($r=4$) at $t=164$. C3 denotes the outward-directed past of a
762: surface just inside the apparent horizon at $t=164$. Thus, moving backwards in time,
763: these curves should asymptote towards the event horizon of the spacetime. These
764: plots suggest that for most of the evolution (at least), the apparent horizon
765: is an excellent approximation to the event horizon.
766: \label{ah_eh}}
767: \end{figure}
768:
769: \section{Conclusions}\label{sec:conclusion}
770: We have performed a preliminary numerical study of the
771: instability of the 5-dimensional black string. Coordinate
772: pathologies prevent us from definitively identifying
773: the final end-state(s) of an unstable black string. This claim is supported
774: by the fact that the code crashes at very nearly the same time at varying
775: resolution, and that curvature invariants remain well behaved throughout
776: the evolution. The former suggests that a numerical instability is not
777: responsible for the crash, while the latter indicates that a physical
778: singularity is probably also not to blame.
779: Despite the premature termination of the simulation, we
780: find evidence that
781: the spacetime evolves towards a configuration that
782: looks like a sequence of black holes connected by thin black
783: strings, and characterized by an expansion of the string direction.
784: Since the spacetime is still
785: fairly dynamical at the time our
786: simulations end, we cannot deduce how close this state is to a final
787: configuration. Nevertheless, the dynamical behavior
788: observed is sufficiently robust for some comments to be made. For instance,
789: the results are not
790: inconsistent with Gregory and Laflamme's conjecture that the
791: solution bifurcates into a sequence of black holes---indeed, we can suggest
792: at least two mechanisms by which this could occur:
793: \begin{enumerate}
794: \item
795: Via a thinning neck that eventually vanishes, if the
796: trend seen in the simulation continues. Note that this would
797: require a) that the proper length of the string continue to grow, in order
798: not to violate area theorems, and b) that the thin string be non-uniform,
799: for otherwise
800: {\em it} would be subject to a further Gregory-Laflamme-like instability.
801: \item
802: Via a sequence of Gregory-Laflamme instabilities, if the thin neck stays
803: ``close'' to a uniform black string, since the neck's length is
804: beyond the critical one for a string with an effective mass computed from
805: the radius of the apparent horizon.
806: In this case, one could
807: envision a ``cascade'' of instabilities leading to the bifurcation.
808: \end{enumerate}
809:
810: We note that either scenario would not necessarily be inconsistent with Horowitz
811: and Maeda's results, should the vanishing of the
812: neck take infinite affine time as measured by local null
813: generators of the horizon. At the same time, a continuation of the
814: observed trend
815: would argue against achieving a stationary solution with a mild dependence
816: on the string dimension (i.e. a small value of $\lambda$), as found in
817: perturbative calculations \cite{gubser}. For then, the rather
818: extreme thinning/bulging that we see must be transient behavior that is
819: ``further'' from the end-state than the perturbed black string was.
820:
821:
822: A more complete exposition of the nature of unstable black string
823: evolution would appear to require coordinate conditions able to adapt to
824: solution features that develop at late times---that is, in a manner that
825: does not introduce severe metric gradients that are not
826: correlated with large gradients in physical quantities.
827: For example, it may help to
828: replace the fixed-lapse slicing with maximal slicing, which
829: enforces that the divergence of the local, spatial volume element
830: be zero. Another, perhaps even more crucial option, would be to
831: introduce a $z$ component to the shift vector that keeps $g_{zz}$
832: close to (or exactly) unity throughout the evolution. These options
833: are currently under investigation.
834:
835: Additionally, it would be interesting to explore a wider range of initial
836: conditions describing ``perturbed'' black strings than that considered
837: here. For example, the imposed
838: $z$-periodicity implies that the equivalent, uncompactified space
839: time consists of identical spherical-black-hole/black-string
840: segments at late (intermediate) times. It would be instructive to
841: see what happens should we break this symmetry, by making $L \gg L_c$,
842: and then introducing some higher-wavelength perturbation similar to
843: $q>1$ in (\ref{gtt}), but with more asymmetry in the initial data (note that
844: this would be more computationally demanding).
845: Finally, it would be very interesting to study the evolution of the
846: solutions recently found by Wiseman~\cite{wiseman}, and mentioned in the
847: introduction.
848: These configurations actually correspond to stationary
849: solutions, and their perturbative stability, or otherwise, is currently not
850: known.
851: Since Wiseman shows that his solutions cannot be the end-states
852: conjectured by Horowitz and Maeda, it is important to
853: understand their behavior, since if they are stable they may well
854: represent physically meaningful states, while if unstable, they may
855: be difficult to attain via dynamical evolution. An interesting observation
856: from our simulations, to the length so far achieved, is that they do not
857: display a conical ``waist'' like those
858: presented in~\cite{wiseman} and further analyzed in~\cite{wisemankol}.
859:
860: \section{Acknowledgments} We gratefully acknowledge support from the
861: following agencies, institutes and grants: NSERC, NSF PHY-0099568, The
862: Canadian Institute for Advanced Research,
863: The Canadian Institute for Theoretical Astrophysics, The Pacific Institute for
864: Mathematical Sciences, The Government of the Basque Country,
865: The Izaak Walton Killam Fund and Caltech's Richard Chase Tolman Fund.
866: Computations were performed on
867: (i) the {\tt vn.physics.ubc.ca} cluster which was funded by the Canadian Foundation for
868: Innovation (CFI) and the BC Knowledge Development Fund; (ii) {\tt LosLobos} at
869: Albuquerque High Performance Computing Center
870: (iii) The high-performance computing facilities within LSU's Center for
871: Applied Information Technology and Learning, which is funded through
872: Louisiana legislative appropriations, and (iv) The {\tt MACI} cluster
873: at the University of Calgary, which is funded by the Universities of Alberta,
874: Calgary, Lethbridge and Manitoba, and by C3.ca, the Netera Alliance, CANARIE,
875: the Alberta Science and Research Authority, and the CFI.
876: We would like to thank G. Horowitz, W.G. Unruh, R.
877: Wald, R. Myers, T. Wiseman and B. Kol for stimulating discussions.
878:
879: \begin{appendix}
880:
881: \section{Initial data solver}\label{app:initialdata}
882: We solve the set of coupled constraint equations
883: (\ref{HC})-(\ref{MC}) via an iterative procedure, where at each sub-step
884: of the iteration we solve a single equation for one of
885: $g_{rr}$, $k_{\theta\theta}$ or $k_{rr}$, assuming that the values
886: of the other variables are known.
887: We iterate this process until the residuals of all the equations are
888: simultaneously below a certain tolerance---a typical value is $10^{-5}$.
889: The overall iteration is initialized using values corresponding to
890: an unperturbed black string solution. We now provide a few more details
891: concerning the solution of each of the constraint equations.
892:
893: The equations are discretized on a uniform grid of points ($x_i,
894: z_j$) with $i=1,...,N_x$, and $j=1,...,N_z$ (recall that, from~(\ref{defx}),
895: the
896: radial coordinate, $r$, is related to $x$ by $r=x/(1-x)$).
897: We first consider the Hamiltonian constraint
898: (\ref{HC}) which, in the coordinate system we have adopted,
899: has the following form:
900: \begin{equation}
901: \label{hc} F_1 \frac{\partial g_{rr}}{\partial x}
902: + F_2 g_{rr} \frac{\partial^2 g_{rr}}{\partial z^2} + F_3 g_{rr}
903: \frac{\partial g_{rr}}{\partial z} + F_4
904: \left(\frac{\partial g_{rr}}{\partial z}\right)^2 + F_5
905: \left(g_{rr}\right)^2 + F_6 g_{rr} =0 \, .
906: \end{equation}
907: Here, the $F_m, \, m = 1, \ldots, 6$, are functions that generally depend
908: on all the metric
909: coefficient and their derivatives {\em except} $g_{rr}$ (and its
910: derivatives). We discretize this equation to second order
911: in the mesh spacing using a difference approximation centered at
912: the points $(x_{i+1/2}, z_j)$.
913: Because of the discretization used and the form of
914: (\ref{hc}), we can solve the resulting set of equations ``line-by-line''
915: in $x$, starting at the inner boundary, $i=1$, which is chosen well
916: within the horizon of the string (typically at $r=M$), and where
917: the boundary values, $[g_{rr}]_{1,j}, \, j = 1,\ldots N_z$, are those
918: corresponding to an unperturbed black string. As we integrate outwards
919: in $x$, the determination of the $i$-th line of unknowns,
920: $[g_{rr}]_{i,j}, \, j = 1,\ldots N_z$, involves the solution of
921: a $N_z$-dimensional, non-linear, cyclic (because of the $z$-periodicity), tridiagonal
922: system that we solve using Newton's method and a cyclic tridiagonal
923: linear solver~\cite{nr}.
924:
925: We now direct attention to the $r$
926: momentum constraint, which, viewed as an equation for $k_{\theta \theta}$,
927: has the form:
928: \begin{equation}
929: G_1 \frac{\partial k_{\theta \theta}}{\partial x} + G_2 k_{\theta
930: \theta} + G_3 =0,
931: \end{equation}
932: where the $G_m, \, m = 1,2,3$ do not depend on $k_{\theta \theta}$ or its
933: derivatives. We note that there is complete decoupling in the $z$-direction
934: in this case; in effect, we have to solve an ODE along each $z={\rm const.}$
935: line. We again discretize using second-order finite difference techniques,
936: fix the boundary values $[k_{\theta\theta}]_{N_x,j}, \, j = 1,\ldots N_z$,
937: at $x=1$ ($i_o$) using the unperturbed
938: black string solution, then solve for the remaining unknowns, marching
939: inwards in $x$.
940:
941: Finally, the $z$ component of the momentum constraint, which fixes
942: $k_{rr}$, has the form:
943: \begin{equation}
944: H_1 \frac{\partial k_{rr}}{\partial z} + H_2 k_{rr} + H_3 =0 \, ;
945: \end{equation}
946: where the $H_m, \, m = 1, 2, 3$ are independent of $k_{rr}$ and its
947: derivatives. This equation is solved analogously to the $r$ momentum
948: constraint, but now using a discretization that is centered at points ($x_i,z_{j+1/2}$).
949: ``Boundary values'', $[k_{rr}]_{i,1}, \, i = 1,\ldots N_x$, are specified along
950: the line $z=z_{{\rm min}}$, again using corresponding values from the
951: black string solution, and the integration proceeds for $j = 2, 3, \ldots N_z$.
952:
953: \section{Finding apparent horizons}\label{app:AH}
954: We use a {\em flow}, or {\em level-set} method to search for
955: apparent horizons within $t={\rm const.}$ spatial slices of
956: the spacetime. We restrict our search to simply connected apparent horizons
957: that are periodic in z. Such an apparent horizon can be described by a curve in
958: the $(r,z)$ plane, which we define to be the level surface $F=0$
959: of the function
960: \begin{equation}\label{fs2}
961: F(r,z) = r - R(z).
962: \end{equation}
963: In other words, the apparent horizon will be given by the curve $r=R(z)$. The
964: apparent horizon
965: is the outermost, marginally trapped surface; hence, we want to
966: find an equation for $R(z)$ such that the outward null expansion,
967: normal to the corresponding surface $F=0$, is zero. To this end,
968: we first construct the unit spatial vector $s^a$, normal to
969: $F={\rm const.}$:
970: \begin{equation}\label{sadef}
971: s^a=\frac{g^{ab} F_{,b}}
972: {\sqrt{g^{cd} F_{,c}F_{,d}}}.
973: \end{equation}
974: Then, using $s^a$ and the $t={\rm const.}$ hypersurface normal vector $n^a$,
975: we can construct future-pointing outgoing($+$) and ingoing($-$) null
976: vectors
977: \begin{equation}\label{null}
978: \ell^a \pm=n^a \pm s^a.
979: \end{equation}
980: The normalization of the null vectors is (arbitrarily)
981: $\ell^a_+ \ell_{-a}=-2$.
982: The outward null expansion $\theta_+$ is then the divergence of $\ell^a_+$
983: projected onto an $F={\rm const.}$ surface:
984: \begin{equation}\label{exp1}
985: \theta_+ = \left(g^{ab}-s^a s^b \right) \nabla_b \ell_{+a}.
986: \end{equation}
987: Substituting expressions (\ref{fs2}) and (\ref{sadef}) into equation (\ref{exp1}),
988: with $\theta_+=0$, provides us with an ordinary differential equation
989: for $R(z)$.
990:
991: Initially we solved equation (\ref{exp1}) via a ``shooting''
992: method---given a guess for $R(0)$, and assuming that
993: $R'(0)=0$ (where a prime denotes differentiation with respect to $z$),
994: we integrate the equation to $z=z_L$, and repeat
995: the process until we find a solution where $R(z_L)=R(z)$ and
996: $R'(z_L)=0$. An efficient sequence of guesses can be
997: generated using a bisection search, as the qualitative behavior of
998: the solution is different depending upon whether the initial guess
999: for $R(0)$ is inside or outside the apparent horizon.
1000:
1001: Since the shooting method is difficult to extend to a parallel
1002: implementation in an efficient way, we opted to use the
1003: following point-wise relaxation method (or flow method) to determine
1004: $R(z)$. We supply an initial guess, $R_0(z)$, for the entire
1005: function $R(z)$, and then iterate the following equation until the
1006: norm of the expansion $\theta_+(z)$ of $F(0)$ is below some
1007: desired threshold (in our runs we have typically set it to
1008: $10^{-3} h$, where $h$ is the basic scale of discretization):
1009: \begin{equation} \label{flow_eqn}
1010: \Delta R_n(z) \equiv R_{n+1}(z) - R_{n}(z) = - \theta_+(z) \Delta
1011: \tau \, .
1012: \end{equation}
1013: Here, $R_{n}(z)$ is the solution after the $n^{th}$ iteration.
1014: Equation~(\ref{flow_eqn}) can be viewed as an {\em explicit}
1015: discretization of a parabolic evolution equation for
1016: $R(z,\tau)$, where $\tau$ is ``time'' and $\Delta \tau$ is the
1017: time-step for the evolution (the parabolic nature of the equation
1018: is evident when $\theta_+$ is expanded in terms of $R(z)$ via
1019: (\ref{exp1})---for brevity we do not give the explicit form of the
1020: equation here). Thus, for stability $\Delta \tau$ must be chosen to
1021: be less than $\Delta z^2$.
1022:
1023: Given a ``reasonable'' initial guess $R_0(z)$, one can see how
1024: iteration of equation (\ref{flow_eqn}) will cause $R_n(z)$ to ``flow''
1025: to the apparent horizon: if $R_n(z)$ is outside of the apparent horizon, then typically the
1026: expansion $\theta_+(z)$ will be positive there, causing
1027: $R_{n+1}(z)$ to decrease towards the apparent horizon, and vice-versa if
1028: $R_n(z)$ is inside the apparent horizon. We use $R(z)=2$ as the
1029: initial guess at $t=0$; after $t=0$ we search for the apparent horizon every $N$
1030: time steps (where $N$ is typically in the range $10-30$), and use
1031: the shape found at the previous search as the initial guess for
1032: the next search. Usually, on the order of tens to thousands of
1033: iterations of (\ref{flow_eqn}) are required to solve for $R(z)$ to
1034: within a level of accuracy such that the approximate solution is
1035: roughly within a mesh spacing of the exact solution (as estimated
1036: in a few specific calculations by solving (\ref{flow_eqn}) close to
1037: machine precision). A single iteration of (\ref{flow_eqn}) can be
1038: computed very rapidly relative to the time taken to compute a
1039: metric-evolution step; however in a parallel environment, if
1040: thousands of iterations are needed on a regular basis (which {\em
1041: is} so at late times during the evolution of an unstable black
1042: string), the apparent horizon finder becomes a slight speed bottleneck in
1043: the code, due to the time it takes to communicate the results of
1044: each iteration
1045: amongst the processors involved.
1046:
1047: \section{Finding (Approximate) Event Horizons}\label{app:null}
1048: Here we describe one method we use, following~\cite{eh}, to locate
1049: approximations to event horizons. This method involves
1050: locating the boundary of the causal past of some $r={\rm const.}$ surface
1051: of the spacetime by following radial null geodesics.
1052:
1053: We write the geodesic equation in Lagrangian form:
1054: \begin{equation}
1055: {\mathcal L} = g_{tt} \left(t^\prime\right)^2
1056: + 2 g_{tr} t^\prime r^\prime
1057: + g_{rr} \left(r^\prime\right)^2
1058: + 2 g_{rz} r^\prime z^\prime
1059: + g_{zz} \left(z^\prime\right)^2,
1060: \end{equation}
1061: where $\lambda$ is the affine parameter along the geodesic, and a
1062: prime denotes differentiation with respect to $\lambda$. Since we are
1063: interested in null trajectories, we set ${\mathcal L}=0$. For radial,
1064: ingoing geodesics, we have $\theta^\prime=\phi^\prime=0$, and thus the geodesic
1065: equations reduce to the set:
1066: %
1067: \begin{eqnarray}
1068: \label{eq:geod}
1069: \dot r & = & \frac{\alpha}{\sqrt{g_{rr}}} - \beta, \\
1070: \dot \lambda & = & \frac{2 \alpha \sqrt{g_{rr}}}{\Pi_r}, \nonumber
1071: \end{eqnarray}
1072: %
1073: where the dot denotes a derivative with respect to coordinate time, and
1074: $\Pi_r={\partial{\mathcal L}}/{\partial r^\prime}$. Then, starting at a certain
1075: value of $r=r_0$, and for each grid point along the $z$ direction, equations
1076: (\ref{eq:geod}) are integrated
1077: backwards in time using a second order Runge-Kutta scheme.
1078:
1079: Following null geodesics along $z={\rm const.}$ lines does not guarantee that we
1080: are tracing the causal past of $r=r_0$, though for the spacetimes studied here,
1081: and the coordinate system used, this should offer a good approximation. Furthermore,
1082: although radial geodesics might not be the best choice, since the event horizon is an
1083: attractor, they will trace it accurately. (For related discussions of
1084: approximate event horizon location in the axisymmetric four-dimensional case,
1085: see~\cite{shapiroteukolskyEH,seidelEVENTHORIZONS,caveny}.)
1086:
1087: \end{appendix}
1088:
1089: \begin{thebibliography}{99}
1090:
1091: \bibitem{chandra} S. Chandrasekhar, {\it The mathematical theory of black holes},
1092: Oxford University Press, Oxford (1983).
1093:
1094: \bibitem{gregorylaflame1}
1095: R. Gregory and R. Laflamme, {\it Phys. Rev. Lett}, {\bf 70} 2837 (1993).
1096:
1097: \bibitem{gregorylaflame2}
1098: R. Gregory and R. Laflamme, {\it Nucl. Phys.}, {\bf B428} 399 (1994).
1099:
1100: \bibitem{hawkingellis}
1101: S. Hawking and G. Ellis, {\it The Large Scale Structure of
1102: spacetime}, Cambridge University Press, Cambridge, (1973).
1103:
1104: \bibitem{hms96}
1105: G.~T.~Horowitz, J.~M.~Maldacena and A.~Strominger,
1106: %``Nonextremal Black Hole Microstates and U-duality'',
1107: {\em Phys.\ Lett.} {\bf B383}, 151 (1996).
1108:
1109: \bibitem{imsy98}
1110: N.~Itzhaki, J.~M.~Maldacena, J.~Sonnenschein and S.~Yankielowicz,
1111: %``Supergravity and the large N limit of theories with sixteen supercharges,''
1112: {\em Phys.\ Rev.} {\bf D58}, 046004 (1998).
1113:
1114: \bibitem{agmoo00}
1115: O.~Aharony, S.~S.~Gubser, J.~M.~Maldacena, H.~Ooguri and Y.~Oz,
1116: %``Large N field theories, string theory and gravity,''
1117: {\em Phys.\ Rept.} {\bf 323}, 183 (2000).
1118:
1119: \bibitem{hv00}
1120: G.~T.~Horowitz and V.~E.~Hubeny,
1121: %``Quasinormal modes of AdS black holes and the approach to thermal equilibrium,''
1122: {\em Phys.\ Rev.} {\bf D62}, 024027 (2000).
1123:
1124: \bibitem{horowitzmaeda}
1125: G. Horowitz and K. Maeda, {\it Phys. Rev. Lett}, {\bf 87} 131301 (2001).
1126:
1127:
1128: \bibitem{gubser} S. Gubser,
1129: %{\it On non-uniform black branes},
1130: CALT-68-2351, CITUSC/01-035, hep-th/0110193, (2001).
1131:
1132: \bibitem{wiseman}
1133: T. Wiseman,
1134: %{\it Static Axisymmetric Vacuum Solutions and Non-Uniform Black Strings},
1135: {\em Class.\ Quant.\ Grav.}, {\bf 20}, 1177 (2003)
1136: %hep-th/0209051 (2002).
1137:
1138: \bibitem{kol}
1139: B. Kol,
1140: %{\it Speculative generalization of black hole uniqueness to higher dimensions},
1141: hep-th/0208056 (2002).
1142:
1143: \bibitem{unruhwald}
1144: W. Unruh and R. Wald. In preparation.
1145:
1146: \bibitem{geddes}
1147: J.~Geddes, {\it Phys. Rev. D}, {\bf 65}, 104015 (2002).
1148:
1149: \bibitem{MTW} C.W. Misner, K.S. Thorne and J.A. Wheeler,
1150: {\em Gravitation}, New York, W.H. Freeman and Company (1973).
1151:
1152: \bibitem{york_78} J.W. York, Jr.,
1153: in {\em Sources of Gravitational Radiation},
1154: ed. L. Smarr, Seattle, Cambridge University Press (1979).
1155:
1156: \bibitem{lehnerreview}
1157: L. Lehner, {\it Class. Quant. Grav.}, {\bf 18}, R25 (2001).
1158:
1159: \bibitem{shinkai}
1160: H.~A.~Shinkai and G.~Yoneda,
1161: {\it Class. Quant. Grav.}, {\bf 19}, 1027 (2002).
1162:
1163: \bibitem{kelly}
1164: B. Kelly, et. al. {\it Phys. Rev. D}, {\bf 64}, 084013 (2001).
1165:
1166: \bibitem{unruh} J. Thornburg, {\it Class. Quant. Grav.}, {\bf 4}, 1119 (1987).
1167:
1168: \bibitem{seidelexcision}
1169: E. Seidel and W.M. Suen,
1170: %{\it Towards a singularity-proof scheme in numerical relativity},
1171: {\em Phys.\ Rev.\ Lett.}, {\bf 69}, 1845 (1992).
1172: %gr-qc/9210016 (1992).
1173:
1174: \bibitem{lehnercadm}
1175: L. Lehner, M. Huq, and D. Garrison, {\it Phys. Rev. D}, {\bf 62}, 084016, (2000).
1176:
1177: \bibitem{cactus}
1178: The CACTUS Computational Toolkit. (http://www.cactuscode.org).
1179:
1180: \bibitem{kreissoliger}
1181: H.O. Kreiss and J. Oliger, {\it Method for the approximate
1182: solution of time dependent problems}, GARP Publication Series,
1183: Geneva (1973).
1184:
1185: \bibitem{eh}
1186: J. Libson, et. al.,{\it Phys. Rev. D.}, {\bf 53}, 4335, (1996).
1187:
1188:
1189: \bibitem{waldah}
1190: R. Wald and S. Iyer, {\it Phys. Rev. D.}, {\bf 44}, 3719, (1991).
1191:
1192: \bibitem{wisemankol}
1193: B.~Kol and T.~Wiseman,
1194: %``Evidence that highly non-uniform black strings have a conical waist,''
1195: arXiv:hep-th/0304070, (2003).
1196:
1197: \bibitem{nr}
1198: W.H. Press, S.A. Teukolsky, W.T. Vetterling and B.P. Flannery,
1199: {\em Numerical Recipes}, Cambridge University Press, (1992).
1200:
1201: \bibitem{shapiroteukolskyEH}
1202: S. Hughes, et. al., {\it Phys. Rev. D.}, {\bf 49}, 4004, (1994).
1203:
1204: \bibitem{seidelEVENTHORIZONS}
1205: J. Masso, et. al.,{\it Phys. Rev. D.}, {\bf 59}, 064015, (1999).
1206:
1207: \bibitem{caveny}
1208: S. A. Caveny and R. A. Matzner,
1209: %{\it Tracking Black Holes in Numerical Relativity },
1210: gr-qc/0303099.
1211:
1212: \end{thebibliography}
1213:
1214:
1215: \end{document}
1216: