1: %
2: % LaTeX template file for
3: % Publications of the Astronomical Society of Australia.
4: % Version 4.0 - 1 May 2004
5: %
6: % The most current version of this file can be found at:
7: % http://www.publish.csiro.au/journals/pasa/template.tex
8: %
9: % Other information on PASA can be found at:
10: % http://www.publish.csiro.au/journals/pasa
11: %
12: % Instructions for submitting to PASA can be found at:
13: % http://www.publish.csiro.au/journals/pasa/authors.html
14: %
15: % VERSIONS:
16: % 3.0 Implements natbib
17: % Thanks to D. Kawata for spotting natbib error
18: %
19: % 4.0 Has changed layout to be closer to final article
20: %
21: \documentclass[a4paper,twoside]{article}
22: %
23: % Baselineskip may be altered if desired.
24: %
25: \baselineskip=2em
26: %
27: % A few definitions.
28: %
29: \def\changemargin#1#2{\list{}{\rightmargin#2\leftmargin#1}\item[]}
30: \let\endchangemargin=\endlist
31: \newcommand{\affil}[1]{$^{\rm #1}$}
32: %
33: % Do not change the page dimensions as these are approximately the size of
34: % the finished article.
35: \textwidth=16.1cm
36: \textheight=23.3 cm
37: \topmargin=-.5 cm
38: \oddsidemargin=0.5cm
39: \evensidemargin=0.5cm
40: \columnsep=0.8cm
41: \renewcommand{\baselinestretch}{.95}
42: %
43: %
44: %%%%%%%%%%%% PAGE HEADERS %%%%%%%%%%%%%%
45: \pagestyle{myheadings}
46: \markboth{\small Publications of the Astronomical Society of Australia}{\small
47: www.publish.csiro.au/journals/pasa}
48: %
49: %
50: %
51: %%%%%%% ADD ADDITIONAL PACKAGES HERE %%%%%%%%%
52: %Citations may be made using the natbib commands \citet{},\citep{} etc.
53: \usepackage[authoryear]{natbib}
54: \bibpunct{(}{)}{;}{a}{}{,}
55: %Use of the graphicx package for figures is recommended, but other well-known
56: %packages, e.g. psfig are also acceptable.
57: \usepackage{graphicx}
58: %
59: %
60: \date{} %Please leave the date blank
61: %
62: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
63: % You may add your own commands here, e.g.
64: \newcommand{\kms}{\mbox{km\,s$^{-1}$}}
65: %
66:
67: %%%%%%%%%%%%% TITLE %%%%%%%%%%%%%%%
68: % PASA titles use initial capitals style, e.g.
69: \title{\large\bf\flushleft No Way Back: Maximizing survival
70: time below the Schwarzschild event horizon}
71: %%%%%%%%%%%% AUTHORS %%%%%%%%%%%%%%
72: \author{\parbox{\textwidth}{\flushleft
73: \vspace{-0.5cm}
74: %
75: % Please indicate only one corresponding author email, as per the following example:
76: {\it Geraint F. Lewis and Juliana Kwan}\\
77: %
78: \vspace{0.4cm}
79: %
80: {\small School of Physics, University of Sydney, NSW 2006} \\
81: {\small Email: gfl,jkwan@physics.usyd.edu.au}
82: }}
83: %
84: %
85: %%%%%%%% DO NOT EDIT FOLLOWING %%%%%%%%%%%%
86: \begin{document}
87: %
88: \twocolumn[
89: \begin{changemargin}{.8cm}{.5cm}
90: \begin{minipage}{.9\textwidth}
91: \vspace{-1cm}
92: \maketitle
93: %
94: %
95: %%%%%%%%%%%%% ABSTRACT %%%%%%%%%%%%%
96: %Abstract of no more than 200 words here.
97: \small{\bf Abstract:} It has long been known that once you cross the event
98: horizon of a black hole, your destiny lies at the central singularity,
99: irrespective of what you do. Furthermore, your demise will occur in a finite
100: amount of proper time. In this paper, the use of rockets in extending the
101: amount of time before the collision with the central singularity is examined.
102: In general, the use of such rockets can increase your remaining time, but only
103: up to a maximum value; this is at odds with the ``more you struggle, the less
104: time you have'' statement that is sometimes discussed in relation to black
105: holes. The derived equations are simple to solve numerically and the
106: framework can be employed as a teaching tool for general relativity.
107:
108: %%%%%%%%%%%%% KEYWORDS %%%%%%%%%%%%%
109: \medskip{\bf Keywords:} black hole physics --- relativity -- methods:
110: numerical --- methods: analytical
111: % Please write all keywords in lower case. PASA uses the
112: % standard list of subject headings adopted by The Astrophysical Journal
113: % and available from http://www.journals.uchicago.edu/ApJ/keywords_text.html.
114: % Keywords are separated by em-dashes, i.e. ---
115:
116: %%%%%%%%DO NOT EDIT%%%%%%%%%%%%
117: \medskip
118: \medskip
119: \end{minipage}
120: \end{changemargin}
121: ]
122: \small
123: %%%%%%%%EDIT FROM HERE%%%%%%%%%%%%
124:
125: \section{Introduction}\label{intro}
126: General relativity is one of the pillars of modern physics, providing an
127: accurate mathematical picture of gravitation and cosmology~\citep[see][for a
128: superb description]{1973grav.book.....M}. While extremely successful, the
129: theory predicts the existence of black holes, completely collapsed massive
130: objects which possess a one-way membrane (the event horizon) through which
131: objects can pass through from the Universe, but not return. The strange
132: properties of these objects has sparked the public
133: imagination~\citep{1994bhtw.book.....T,1995bhu..book.....N} and are the staple
134: of most undergraduate courses on general relativity.
135:
136: In this article, the question of the journey within the event horizon is
137: examined, especially with regards to attempts to prolong, through the use of
138: powerful rockets, the time to the inevitable collision with the central
139: singularity at $r=0$. While touched upon in many texts, the discussion of
140: their use in the vicinity of black holes is not common. Hence this article is
141: a pedagogical study of the use of coordinates and physical acceleration in
142: general relativity. Furthermore, it aims to clear up a few black hole myths,
143: especially those that appear on authoritative Internet
144: websites~\footnote{While the authors acknowledges that the Internet is not the
145: ultimate font of knowledge, anyone who has marked a few undergraduate essays
146: will know that many students see it as their only source of knowledge.}. In
147: Section~\ref{hist} a little history is presented, while Section~\ref{setup}
148: outlines the approach taken. The results of this study appear in
149: Section~\ref{results} and the conclusions in Section~\ref{conclusions}.
150:
151: \begin{figure*}
152: \begin{center}
153: \includegraphics[scale=0.67, angle=-90]{fig1.ps}
154: \caption{The left hand panel presents several free fall paths into a black
155: hole. The paths begin at $3.0m$ (black), $2.5m$ (red) $2.1m$ (green) and
156: $2.00000001m$ (blue). The solid curves represent the path in terms of the
157: proper time of the faller, while the dashed path is with respect to the
158: coordinate time in Eddington-Finkelstein coordinates. The right hand panel
159: presents the conserved quantity $e$ (Equation~\ref{kill})}\label{fig1}
160: \end{center}
161: \end{figure*}
162:
163:
164: \section{A little history}\label{hist}
165: It has been ninety years since Schwarzschild presented the first exact
166: solution to the field equations of general relativity \citep{Schwarz1916}.
167: Representing the spacetime curvature outside of a spherical mass distribution,
168: the existence of singularities in the solution led to several confusing
169: problems. Importantly, the coordinate time is seen to diverge as an object
170: falling in this spacetime approaches the Schwarzschild radius ($r=2m$, in
171: units where $G=c=1$ and where $m$ is the mass of the black hole), with the
172: conclusion that the entire history of the Universe can pass before anything
173: actually falls to this radius. Paradoxically, the proper time as experienced
174: by the falling object is finite through $r=2m$ and the faller reaches $r=0$ in
175: finite time.
176:
177: A reformulation of the Schwarzschild solution in free-falling coordinates
178: revealed the Schwarzschild radius to be an event horizon, a boundary which can
179: only be crossed from $r>2m$, but not from $r<2m$, leading to notion of
180: complete gravitational collapse and the formation of black holes
181: \citep{painleve}. However, even with these advances, the singular state of the
182: Schwarzschild solution at $r=2m$ led even the most famous relativist to
183: suggest that black holes cannot form~\citep{Ein1939}. The resolution was
184: ultimately provided by \citet{Fink1958} who derived a coordinate
185: transformation of the Schwarzschild solution which made it finite at $r=2m$;
186: this was, however, a rediscovery of the earlier work by \citet{Edd1924} who
187: apparently did not realize its significance~\footnote{It is more astounding
188: that in his analysis, \citet{Edd1924} explicitly considered outgoing light
189: rays which, in his transformed coordinates, clearly crossed the event
190: horizon from the inside to the outside. While he did not note it, Eddington
191: had uncovered the white hole Schwarzschild solution.}. With this
192: transformation the true nature of the Schwarzschild radius was revealed,
193: acting as a one-way membrane between the Universe and inner region of the
194: black hole. Surprisingly, the analysis of \citet{Fink1958} also possesses a
195: time reversed black hole solution, a white hole in which the one-way membrane
196: is reversed.
197:
198: As discussed in many texts, the transformation to Eddington-Finkelstein
199: coordinates clearly reveals the ultimate fate of an infalling observer. Now
200: crossing the event horizon in a finite coordinate time, the future light cones
201: for all massive explorers are tilted over such that there is no way back and
202: the future ultimately lies at the central singularity. But after crossing the
203: horizon, how long does the intrepid explorer have until this happens, and what
204: can they do to maximize their survival time? For a free-falling path, the
205: calculation of the proper time experienced by the explorer is a question found
206: in graduate texts~\citep[e.g. see problem 12-14 in][]{2003gieg.book.....H} and
207: it is straightforward to show that the maximum time that can be experienced
208: below the event horizon is
209: \begin{equation}
210: \tau = \pi m
211: \end{equation}
212: For a stellar mass black hole, this will be a fraction of a second, but for a
213: supermassive black hole, this may be hours. As will be shown later, this
214: maximum time applies to a faller who drops from rest at the event horizon and
215: any one who starts falling from above the event horizon and free falls into
216: the hole will experience less proper time on the journey from the event
217: horizon to the singularity.
218:
219: \section{Setting Up the Problem}\label{setup}
220: In this paper, only purely radial motion will be considered and the faller
221: will be assumed to be impervious to the significant inertial and tidal forces
222: it will suffer on its journey.
223:
224: \subsection{Eddington-Finkelstein Metric}\label{Edd}
225:
226: In considering a radial journey across the event horizon, the advanced
227: Eddington-Finkelstein coordinates will be employed. With this, the
228: Schwarzschild solution is represented by the invariant interval of the
229: form~\footnote{ There is more than one representation of the
230: Eddington-Finkelstein metric for the Schwarzschild solution, and often it is
231: written in terms of an advanced time parameter. However, as this parameter
232: is null, the metric is often recast in terms of a new time-like parameter,
233: resulting in the metric given above [see Chapter 11.5 in
234: \citet{2005gere.book.....H}]; this is explicitly the form of the metric
235: investigated by \citet{Edd1924} and \citet{Fink1958}. }
236: \begin{equation}
237: ds^2 = -\left( 1 - \frac{2m}{r} \right) dt^2 + \frac{4m}{r} dt dr
238: + \left( 1 + \frac{2m}{r} \right) dr^2 + r^2 d\Omega^2
239: \label{eddfink}
240: \end{equation}
241: As noted previously, in this form the interval is non-singular at the event
242: horizon $(r = 2m)$.
243:
244: \begin{figure*}
245: \begin{center}
246: \includegraphics[scale=0.67, angle=-90]{fig2.ps}
247: \caption{As in Figure~\ref{fig1}, for observers falling from $r=3m$. Here, the
248: black line represents a free faller, while the red, green and blue
249: represents a rocketeer accelerating outwards $a=0.5, 2.5$ and $5.0$
250: respectively.}\label{fig2}
251: \end{center}
252: \end{figure*}
253:
254:
255: \subsection{4-velocity and 4-acceleration}\label{4velacc}
256: The majority of texts on general relativity consider free fall motion through
257: spacetime, with no acceleration terms due to non-gravitational forces. Such
258: free fall paths are governed by the well-known geodesic equation which
259: parameterizes the coordinates, $x^\alpha$ of a massive object in terms of its
260: proper time, $\tau$,
261: \begin{equation}
262: x^\alpha = (t(\tau),r(\tau),\theta(\tau),\phi(\tau))
263: \end{equation}
264: From this, it is simple to define a 4-velocity, $u^\alpha$, of the form
265: \begin{equation}
266: u^\alpha = \frac{dx^\alpha}{d\tau} = \left(\frac{dt}{d\tau},\frac{dr}{d\tau},
267: \frac{d\theta}{d\tau},\frac{d\phi}{d\tau}\right)
268: \end{equation}
269: If the massive body undergoes a 4-acceleration, $a^\alpha$, due to a force,
270: the equation of motion can be written as
271: \begin{equation}
272: a^\alpha = \frac{du^\alpha}{d\tau} + \Gamma^\alpha_{\beta\gamma} u^\beta
273: u^\gamma
274: \label{geo}
275: \end{equation}
276: where $\Gamma^\alpha_{\beta\gamma}$ are the Christoffel symbols or affine
277: connections; clearly, if the 4-acceleration is zero, the standard geodesic
278: equation is recovered. The required Christoffel symbols are simply calculated
279: from the Eddington-Finkelstein metric using {\tt GRTensor}\footnote{\tt
280: grtensor.phy.queensu.ca/} in Mathematica. The non-zero components needed
281: for this study are
282: \begin{equation}
283: \begin{array}{ll}
284: \Gamma^t_{tt} = \frac{2m^2}{r^3} &
285: \Gamma^t_{rr} = \frac{2m(m+r)}{r^3} \\
286: & \\
287: \Gamma^r_{tt} = \frac{m(r-2m)}{r^3} &
288: \Gamma^r_{rr} = \frac{-m(2m+r)}{r^3} \\
289: & \\
290: \Gamma^t_{tr}=\Gamma^t_{rt} = \frac{m(2m+r)}{r^3} &
291: \Gamma^r_{tr}=\Gamma^r_{rt} = \frac{-2m^2}{r^3} \\
292: \end{array}
293: \end{equation}
294: The
295: path of an accelerated object is also constrained through the normalization of
296: the 4-velocity of a massive particle
297: \begin{equation}
298: {\bf u} \cdot {\bf u} = g_{\alpha\beta} u^\alpha u^\beta = -1
299: \label{4vel}
300: \end{equation}
301: and its orthogonality with the 4-acceleration
302: \begin{equation}
303: {\bf a} \cdot {\bf u} = g_{\alpha\beta} a^\alpha u^\beta = 0
304: \label{4norm}
305: \end{equation}
306: where $g_{\alpha\beta}$ are the components of the metric (Equation ~\ref{eddfink}).
307: The final constraining equation is the normalization of the 4-acceleration
308: \begin{equation}
309: {\bf a} \cdot {\bf a} = g_{\alpha\beta} a^\alpha a^\beta = a^2
310: \label{4acc}
311: \end{equation}
312: where $a$ is the magnitude of the acceleration. Note that this also
313: represents the magnitude of the acceleration as experienced by our faller due
314: to the presence of the rockets.
315:
316: \begin{figure*}
317: \begin{center}
318: \includegraphics[scale=0.67, angle=-90]{fig3.ps}
319: \caption{As in Figure~\ref{fig1}, except each faller undergoes an outward
320: acceleration of $a=0.5$ once inside the event horizon. }\label{fig3}
321: \end{center}
322: \end{figure*}
323:
324: \subsection{Hyperbolic Motion}\label{hyperbolic}
325: In an insightful paper, \citet{1960PhRv..119.2082R} demonstrated that all
326: bodies undergoing constant acceleration undertake hyperbolic motion; while
327: this result is well known in the framework of special relativity, this paper
328: was the first to determine that accelerated bodies execute hyperbolic motion
329: in the curved spacetime of general relativity \citep[see
330: also][]{1969PhRv..185.1662G,1971PhRvD...3.1035K}. Through an examination of
331: the geometry of motion, \citet{1960PhRv..119.2082R} showed that the components
332: of the 4-velocity and 4-acceleration can be given in terms of two other
333: tensors, $M^\alpha$ and $L^\alpha$, such that
334: \begin{eqnarray}
335: u^\alpha = & (cosh\ a\tau)L^\alpha + (sinh\ a\tau) M^\alpha \nonumber\\
336: a^\alpha = & a[(sinh\ a\tau)L^\alpha + (cosh\ a\tau) M^\alpha]
337: \end{eqnarray}
338: where $a$ is the magnitude of the acceleration (Equation~\ref{4acc}) and
339: $\tau$ is the proper time as measured by the accelerated body. The tensors
340: $L^\alpha$ and $M^\alpha$, are orthogonal unit-vectors, being time-like and
341: space-like respectively. Operationally, these tensors are parallel-propagated
342: along the path of the accelerated motion, such that
343: \begin{eqnarray}
344: \frac{dL^\alpha}{d\tau} + \Gamma^\alpha_{\beta\gamma} L^\beta
345: u^\gamma = 0 \nonumber \\
346: \frac{dM^\alpha}{d\tau} + \Gamma^\alpha_{\beta\gamma} M^\beta
347: u^\gamma = 0
348: \label{nogeo}
349: \end{eqnarray}
350: and the initial conditions can be set by noting that at $\tau=0$, then
351: $L^\alpha = u^\alpha$ and $M^\alpha = a^\alpha / a$. Hence, given a fixed
352: magnitude of acceleration, $a$, the normalization equations in the previous
353: section can be used to determine the components of the 4-acceleration $a^t$
354: and $a^r$. With this, the equations of motion can be derived from
355: Equation~\ref{nogeo} and the resulting coupled differential equations were
356: integrated with {\tt odepack}\footnote{\tt www.llnl.gov/CASC/odepack/}.
357:
358: \subsection{Killing Vectors and Conserved Quantities}\label{killing}
359: In treating physical problems, conserved quantities are often employed to ease
360: the understanding of the solutions. In general relativity, these are provided
361: by Killing vectors. Simply put, a Killing vector `points' in a direction along
362: which the metric does not change. For a given Killing vector, $\xi^\alpha$, a
363: conserved quantity can be found for an object that moves along a geodesic to
364: be
365: \begin{equation}
366: e = {\bf \xi}\cdot{\bf u} = g_{\alpha\beta} \xi^\alpha u^\beta
367: \end{equation}
368: Clearly, the components of the Eddington-Finkelstein representation of the
369: Schwarzschild solution (Equation~\ref{eddfink}) are independent of the $t$
370: coordinate (i.e. a translation in this coordinate leaves the metric the same)
371: and its associated Killing vector is given by $\xi^\alpha=(1,0,0,0)$ and the
372: resultant conserved quantity is
373: \begin{equation}
374: e = g_{tt} u^t + g_{tr} u^r
375: \label{kill}
376: \end{equation}
377: It must be remembered that this quantity is conserved along geodesics and so
378: only for freely-falling objects. For objects undergoing acceleration (e.g. due to
379: rockets), this quantity is not conserved. This has significant implications for
380: maximizing the proper time below the event horizon.
381:
382: With the above definition of the conserved quantity related to the Killing
383: vector, as well as the 4-velocity and 4-acceleration normalization and
384: orthogonality, a little algebra reveals that for an acceleration of magnitude
385: $a$, then
386: \begin{equation}
387: a^r = a \frac{u^r e}{\sqrt{e^2 + g_{tt}}}
388: \end{equation}
389: and
390: \begin{equation}
391: a^t = \frac{(1 + u^t e)}{u^r e} a^r
392: \end{equation}
393: It is important to remember that in the presence of a non-zero acceleration,
394: the quantity $e$ is no longer conserved.
395:
396: \section{Results}\label{results}
397:
398: \begin{figure*}
399: \begin{center}
400: \includegraphics[scale=0.67, angle=-90]{fig4.ps}
401: \caption{As in Figure~\ref{fig1}, with the black line representing a free
402: faller from $r=3m$. The other lines correspond to an observer who free falls
403: to the event horizon and then fires their rocket with $a=2$. For the red line,
404: the faller fires their rocket all the way to the singularity, while the dark
405: blue, light blue and green turn off their rocket when $e=0.3$, $e=-0.3$ and
406: $e=0$ respectively. An examination of the proper time in the left-hand panel
407: reveals that it the path that settles on $e=0$ that possesses the longest
408: proper time.
409: }\label{fig4}
410: \end{center}
411: \end{figure*}
412:
413: \subsection{Analytic checks}\label{analytic}
414: Before considering the influence of the rocket, it is important to check the
415: computational solutions with comparison to analytic results for freely falling
416: objects. Assuming the faller begins from rest beyond the event horizon at a
417: radius $r_s$, so $u^r(r_s)=0$, then the conserved quantity given by the
418: Killing vector (Equation~\ref{kill}) is
419: \begin{equation}
420: e = g_{tt} u^t = -\sqrt{ 1 - \frac{2m}{r_s}}
421: \label{cons}
422: \end{equation}
423: where $u^t$ at $r_s$ is determined from the normalization of the 4-velocity
424: (Equation~\ref{4vel}). Clearly, if the faller starts from $r_s=\infty$ then
425: $e=-1$ and, conversely, if the faller drops from rest at the event horizon,
426: $(r_s=2m)$, then $e=0$. As noted previously, the free fall journey from rest
427: at a particular radius to the central singularity is discussed in many text
428: books and will not be reproduced here, but it can be shown that the proper
429: time as measured by the faller is given by
430: \begin{equation}
431: \tau_{max} = \frac{\pi}{2\sqrt{2m}} r_s^{\frac{3}{2}}
432: \label{maxtime}
433: \end{equation}
434: \citep[e.g. see problem 12-5 in][]{2003gieg.book.....H}. Note this is the
435: proper
436: time for the entire journey. The time spent on the portion of the trip between
437: the event horizon and central singularity is given by
438: \begin{equation}
439: \tau = \left\{\frac{1}{\sqrt{2}} \left[\frac{r_s}
440: {m}\right]^{\frac{3}{2}} atan\left[\sqrt{\frac{2m}{r_s-2m}}\right]
441: - \frac{\sqrt{ r_s ( r_s - 2m )}}{m} \right\} m
442: \label{time}
443: \end{equation}
444: For all $r_s>2m$, the proper time experienced by the faller between the event
445: horizon and the singularity is less than Equation~\ref{maxtime}. Conversely,
446: the minimum time that can be experienced by a free-faller (found by taking the
447: limit of $r_s\rightarrow\infty$) is
448: \begin{equation}
449: \tau_{min} = \frac{4}{3} m
450: \end{equation}
451:
452: Figure~\ref{fig1} presents the results of the numerical integration of
453: Equation~\ref{nogeo}, assuming the rockets are not used and so the
454: acceleration terms are zero. For this example, four paths are examined,
455: differing only in the radial coordinate from which they are dropped from rest;
456: these are $3.0m$ (black), $2.5m$ (red) $2.1m$ (green) and $2.00000001m$
457: (blue). Note, as the normalization of the 4-velocity diverges for an object at
458: rest at the event horizon, it is not possible to numerically integrate these
459: equations with the initial condition of $r_s=2m$. In comparing the numerical
460: results for the proper time below the even horizon with the analytic
461: predictions (Equation~\ref{time}), the maximum fractional error is found to be
462: $\sim0.005\%$. Similarly, the fractional error in the conserved quantity
463: (Equation~\ref{kill}) is of a similar order over the journey to the singularity.
464: \subsection{Turning on the rocket}\label{rocket}
465: For the purposes of this study, it is assumed that the faller begins from rest
466: at some distance beyond the black hole, free falling to the event horizon.
467: Once across the horizon, the rocket is ignited. Figure~\ref{fig2} presents the
468: case where such an object is dropped from rest at $r=3m$, with the black curve
469: representing a free falling path (again, the solid line represents the curve
470: with respect to proper time, while the dotted line is that for coordinate
471: time). For the red curve, the rocketeer ignites the rocket as they pass $r=2m$
472: and undergoes a constant, outward acceleration of $a=0.5$, while the green and
473: blue lines suffer an acceleration of $a=2.5$ and $a=5$ respectively. Looking
474: at the left hand panel, it is clear that the use of a rocket can increase
475: the proper time of the faller beyond that expected for a purely free fall path
476: (e.g. the red line). However, it is also apparent there is a limit to the
477: increased proper time through firing the rocket as the more extreme
478: accelerations (green and blue line) experience less proper time than the free
479: calling observer on their journey to the singularity.
480:
481: \begin{figure*}
482: \begin{center}
483: \includegraphics[scale=0.67, angle=-90]{figx.ps}
484: \caption{As in Figure~\ref{fig1}, with the black line representing a free
485: faller, while the red line represents a rocketeer who, once across the event
486: horizon, accelerates inwards for a short while and then accelerates outwards.
487: The amount of acceleration is tuned so that both the free faller and the
488: rocketeer arrive at the central singularity at the same coordinate time
489: (dotted paths). As revealed by the solid paths, the free faller
490: experiences the greater proper time in the journey below the event
491: horizon.}\label{figx}
492: \end{center}
493: \end{figure*}
494:
495: An examination of the conserved quantity from the Killing vector, $e$ in the
496: right hand panel tells an interesting story; free falling from rest outside
497: the event horizon, all of the fallers have the same value of $e$, but once the
498: rocket is fired inside the event horizon, the firing of the rocket increases
499: the value of $e$, and, moreover, the change appears to be linear. In
500: examining this, it is straightforward to show, through a little algebra,
501: that\footnote{An examination of a uniformally accelerating observer in special
502: relativity displays the same relationship.}
503: \begin{equation}
504: \frac{de}{dr} = \frac{g_{tt} a^t + g_{tr} a^r}{u^r} = -a
505: \label{change}
506: \end{equation}
507: Figure~\ref{fig3} shows the free fall paths of observers from several
508: different radii to the event horizon. Once within the horizon, each observer
509: fires their rocket with the same acceleration ($a=0.25$) and continues their
510: journey to the central singularity. As expected from Equation~\ref{change},
511: the quantity $e$ is conserved along the free fall path, but once the rocket
512: is fired $e$ changes linearly with the radial coordinate.
513:
514: Armed with this knowledge, what should an observer who has fallen from outside
515: the event horizon do to maximize they survival time below the event horizon,
516: if they have at their disposal a rocket that can produce an acceleration $a$?
517: As noted earlier, the longest free fall time below the event horizon occurs
518: for an observer who falls from rest at $r=2m$ (with $e=0$) and any attempt at
519: accelerated motion for this observer will only diminish the proper time (this
520: is discussed in more detail in the next section). Hence, if the observer
521: starts from beyond the event horizon with any non-zero value of $e$, the best
522: they can do is fire their rocket until $e$ equals zero and then turn the
523: rocket off and coast on the $e=0$ geodesic to the central singularity. This is
524: illustrated in Figure~\ref{fig4} for several observers who falls from rest at
525: $r=3m$ to the event horizon. Once within the horizon, one rocketeer (black
526: curve) continues their free fall path to the singularity, while the others
527: fire their rockets (with $a=2$). The red path is that of the observer who
528: continues to fire their rocket all the way down, while the light blue, dark
529: blue and green cease firing when $e=-0.3$, $e=0.3$ and $e=0$ respectively. An
530: examination of the left-hand panel of this figure shows that, in terms of
531: coordinate time, the act of firing the rocket delays the collision with the
532: central singularity. However, the time as measured by each observer displays a
533: quite different behaviour; firing the rocket in this circumstance increases
534: the proper time between the horizon and the singularity. However, it is clear
535: that the observer who settles on the path with $e=0$ experiences the greatest
536: proper time, with those that burn their rocket for shorter or longer periods
537: experiencing shorter proper times.
538:
539: \subsection{Clearing up a mythconception}\label{myths}
540: As noted previously, black holes have fired the imagination of the general
541: public and many websites can be found that are dedicated to discussing their
542: strange properties. However, some authoritative websites carry statements
543: like the following\footnote{{\tt
544: cosmology.berkeley.edu/Education/BHfaq.html}}
545: \begin{quote}
546: A consequence of this is that a pilot in a powerful rocket ship that had just
547: crossed the event horizon who tried to accelerate away from the singularity
548: would reach it sooner in his frame, since geodesics (unaccelerated paths) are
549: paths that maximize proper time
550: \end{quote}
551: The results of this study show that this clearly is not the case; anyone who
552: falls through the event horizon should fire their rockets to maximize the time
553: they have left before impacting the central singularity. In dropping from
554: rest at the event horizon, the firing of a rocket does not extend the time
555: left, it only diminishes it.
556:
557: While the quote is ambiguous about the initial conditions for the faller, it
558: appears that the error lies in the assumption that the impact onto the central
559: singularity is the same event for the free faller and the rocketeer; if they
560: were then the above statement would be correct and the free faller would
561: experience the maximal proper time. As an example of this, consider
562: Figure~\ref{figx}. Again, the two fallers start from rest and drop towards the
563: event horizon. After crossing the horizon, one continues the free fall path
564: towards the central singularity while the second accelerates inwards for a
565: short while and then swings their rocket round to accelerate outwards such
566: that both fallers arrive at the central singularity at the same coordinate
567: time (the dotted path). In considering the two paths connecting the two
568: identical events, clearly the proper time as measured by the free faller below
569: the event horizon is greater than that for the rocketeer.
570:
571: \section{Conclusions}\label{conclusions}
572: Black holes remain amongst the most studied theoretical consequences of
573: general relativity, although standard texts say little about the use of
574: rockets once you are below the event horizon. This paper has considered this
575: very scenario, showing that a rocketeer can enhance their survival time by
576: firing a rocket once across the event horizon. However, the rocketeer is
577: still doomed to impact on the central singularity in less than the maximal
578: free fall time between the event horizon and the centre.
579:
580: Additionally, this paper has considered an apparent confusion on the use of
581: rockets below the event horizon which suggest they hasten a fallers demise.
582: This is at odds with this study which shows that rockets can increase survival
583: time for virtually all fallers.
584:
585: Finally, it should be remembered that ingoing light rays in
586: Eddington-Finkelstein coordinates travel at 45$^o$. A simple examination of
587: Figure~\ref{fig4} reveals that something quite interesting; while the
588: constantly accelerating observing within the event horizon (red line)
589: experiences less proper time in their fall to the singularity than the path
590: that settles on $e=0$, an examination of the paths in coordinate time shows
591: that the constantly accelerating observers sees a longer period of time pass
592: in the outside universe than the path on $e=0$. A more detailed study of this
593: effect will be the subject of a future contribution.
594:
595: \section*{Acknowledgments} %If needed
596: James Hartle is thanked for his interesting discussions on the nature of black
597: holes. GFL thanks Matthew Francis and Richard Lane for putting up with his
598: bursting into their office and lecturing them on his eureka moments. The
599: authors would appreciate notification of the use of any material in this
600: article for teaching purposes.
601:
602: \begin{thebibliography}{}
603: \bibitem[Chandrasekhar(1983)Chandrasekhar]{1983mtbh.book.....C}
604: Chandrasekhar, S.\ 1983,
605: Research supported by NSF.~Oxford/New York, Clarendon Press/Oxford
606: University Press (International Series of Monographs on Physics.~Volume
607: 69).
608:
609: \bibitem[Eddington(1924)Eddington]{Edd1924}
610: Eddington, A. S.\ 1924, Nature, 113, 192
611:
612: \bibitem[Einstein(1939)Einstein]{Ein1939}
613: Einstein, A.\ 1939, Annals of Mathematics, 40, 922
614:
615: \bibitem[Finkelstein(1958)Finkelstein]{Fink1958}
616: Finkelstein, D.\ 1958, Physical Review 110, 965
617:
618: \bibitem[Gautreau(1969)]{1969PhRv..185.1662G} Gautreau, R.\ 1969, Physical
619: Review , 185, 1662
620:
621: \bibitem[\protect\citeauthoryear{Hartle}{2003}]{2003gieg.book.....H} Hartle
622: J.~B., 2003, Addison Wesley
623:
624: \bibitem[\protect\citeauthoryear{Hobson, Efstathiou, \& Lasenby}{2005}]{2005gere.book.....H}
625: Hobson M.~P., Efstathiou G.~P., Lasenby A.~N., 2005,
626: `General Relativity: An Introduction for Physicists' (Cambridge University
627: Press: Cambridge, UK)
628:
629: \bibitem[Karlov \& Rindler(1971)]{1971PhRvD...3.1035K} Karlov, L., \&
630: Rindler, W.\ 1971, Ph. Rev. D, 3, 1035
631:
632: \bibitem[Misner et al.(1973)Misner et al.]{1973grav.book.....M}
633: Misner, C.~W., Thorne, K.~S., \& Wheeler, J.~A.\ 1973,
634: Gravitation, San Francisco: W.H.~Freeman and Co.
635:
636: \bibitem[\protect\citeauthoryear{Novikov}{1995}]{1995bhu..book.....N}
637: Novikov I.~D., 1995, Cambridge University Press
638:
639: \bibitem[Painlev\'{e}(1921)Painlev\'{e}]{painleve}
640: Painlev\'{e}, P.\ 1921 C. R. Acad. Sci., 173, 677
641:
642: \bibitem[Pais(1982)Pais]{1982sils.book.....P} Pais, A.\ 1982,
643: Subtle is the Lord, Oxford: University Press
644:
645: \bibitem[Rindler(1960)]{1960PhRv..119.2082R}
646: Rindler W., 1960, PhRv, 119, 2082
647:
648: \bibitem[Schwarzschild(1916)Schwarzschild]{Schwarz1916}
649: Schwarzschild, K.\ 1916
650: Sitzungsberichte der Königlich Preussischen Akademie der Wissenschaften 1, 189
651:
652: \bibitem[\protect\citeauthoryear{Thorne}{1994}]{1994bhtw.book.....T} Thorne
653: K.~S., 1994, Black holes and time warps: Einstein's outrageous legacy,
654: .W. Norton and London, Picador
655:
656: \end{thebibliography}
657:
658: %\end{multicols}
659: \end{document}
660:
661:
662: