hep-th0401209/v2.tex
1: % On the Resolution of the Time-Like Singularities in Reissner-Nordstr\"{o}m
2: % and Negative-Mass Schwarzschild
3: % Amit Giveon, Barak Kol, Amos Ori, Amit Sever
4: 
5: \documentclass{JHEP3}
6: \usepackage{epsfig}
7: %\usepackage{epsfig,amsmath}
8: %\documentclass{article}
9: 
10: \newcommand{\be}{\begin{equation}}
11: \newcommand{\ee}{\end{equation}}
12: \newcommand{\bea}{\begin{eqnarray}}
13: \newcommand{\eea}{\end{eqnarray}}
14: \newcommand{\IR}{\mathbb{R}} \newcommand{\IC}{\mathbb{C}}
15: \newcommand{\IN}{\mathbb{N}}
16: \newcommand{\IQ}{\mathbb{Q}}
17: \def\IZ{\relax\ifmmode\hbox{Z\kern-.4em Z}\else{Z\kern-.4em Z}\fi}
18: \newcommand{\IS}{{\bf S}}
19: \newcommand{\SLtz}{SL(2,{\bf Z})}
20: \newcommand{\SLt}[1]{SL(2,#1)}
21: \newcommand{\pq}{$(p,q)$}
22: \newcommand{\Slash}{\hspace{-2.2mm} /}
23: \newcommand{\non}{\nonumber \\}
24: \newcommand{\sfv}{${\bf S}^5~$}
25: \newcommand{\cpt}{${\bf CP}^2~$}
26: \def\half{{1 \over 2}} \def\quart{{1 \over 4}}
27: \def\tr{{\rm tr}}
28: \def\del{{\partial}}
29: \def\room{~\rule[-2mm]{0mm}{8mm}}
30: 
31: \newcommand{\h}[1]{{\hat #1}} \newcommand{\tl}[1]{{\tilde #1}}
32: \def\bw{\bar{w}}
33: \def\th{{\tilde h}}
34: \def\hh{{\hat h}}
35: \def\ba{{\bar 1}}
36: \def\bb{{\bar 2}}
37: \def\bc{{\bar 3}}
38: \def\bz{{\bar z}} \def\bh{{\bar h}}
39: \def\bx{{\bar x}}
40: \def\bi{{\bar i}}
41: \def\bj{{\bar j}} \def\bk{{\bar k}}
42: \def\bl{{\bar l}} \def\bm{{\bar m}}
43: \def\bl{{\bar n}} \def\bm{{\bar p}}
44: \def\bl{{\bar q}} \def\bm{{\bar r}}
45: \def\cm{{\cal M}} \def\cn{{\cal N}}
46: \def\cl{{\cal L}} \def\co{{\cal O}}
47: \def\tlr{\tilde{r}}
48: \def\cn{{\cal N}} \def\cM{{\cal M}} \def\Mc{{\cal M}_c}
49: \def\hl{\hat l}
50: 
51: %Greek
52: \def\al{\alpha} \def\bt{\beta}
53: \def\gm{\gamma} \def\dl{\delta} \def\eps{\epsilon}
54: \def\hal{{\hat \alpha}} \def\hbt{{\hat \beta}}
55: \def\hgm{{\hat \gamma}} \def\hdl{{\hat \delta}}
56: \def\trho{{\tilde \rho}} \def\hrho{{\hat \rho}}
57: \def\hmu{{\hat \mu}}
58: 
59: \newcommand{\sbsection}[1]{\vspace{.5cm} \noindent {\it #1}}
60: \def\room{~\rule[-2mm]{0mm}{8mm}}
61: \def\scri{{\cal I}}
62: \def\phiom{\phi_\omega}
63: \def\RN{Reissner-Nordstr\"{o}m }
64: \def\Schw{Schwarzschild }
65: \def\KS{Kruskal-Szekeres }
66: 
67: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
68: \def\({\left(}
69: \def\){\right)}
70: %\def\[{\left[}
71: %\def\]{\right]}
72: 
73: \def\br{\left[}
74: \def\kt{\right]}
75: 
76: \def\bra#1{\left\langle #1\right|}
77: \def\ket#1{\left| #1\right\rangle}
78: \def\vev#1{\left\langle #1 \right\rangle}
79: \def\det{{\rm det}}
80: \def\tr{{\rm tr}}
81: \def\mod{{\rm mod}}
82: \def \sinh{{\rm sinh}}
83: \def \cosh{{\rm cosh}}
84: \def \tanh{{\rm tanh}}
85: \def \coth{{\rm coth}}
86: \def \sgn{{\rm sgn}}
87: \def\det{{\rm det}}
88: \def\exp{{\rm exp}}
89: \def\sh{{\rm sinh}}
90: \def\ch{{\rm cosh}}
91: \def \th{{\rm tanh}}
92: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
93: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
94: %title page
95: \preprint{{\tt hep-th/0401209}}
96: \title{On the Resolution of the Time-Like Singularities in
97: Reissner-Nordstr\"{o}m and Negative-Mass Schwarzschild}
98: 
99: \author{Amit Giveon\footnotemark[1], Barak Kol\footnotemark[1],
100:  Amos Ori\footnotemark[2], Amit Sever\footnotemark[1] \\
101: 
102: \footnotemark[1]
103:  Racah Institute of Physics \\
104:  The Hebrew University \\
105:  Jerusalem 91904 \\
106:  Israel\\
107: {\tt giveon, barak\_kol, asever @phys.huji.ac.il}
108: \\
109: 
110: \footnotemark[2] Department of Physics\\
111:   Technion-Israel Institute of Technology\\
112:   Haifa 32000\\
113:   Israel \\
114:   \email{amos@physics.technion.ac.il} }
115: 
116: 
117: \abstract{Certain time-like singularities are shown to be resolved
118: already in classical General Relativity once one passes from
119: particle probes to scalar waves. The time evolution can be defined
120: uniquely and some general conditions for that are formulated. The
121: \RN singularity allows for communication through the singularity
122: and can be termed ``beam splitter'' since the transmission
123: probability of a suitably prepared high energy wave packet is
124: 25\%. The high frequency dependence of the cross section is
125: $\omega^{-4/3}$. However, smooth geometries arbitrarily close to
126: the singular one require a finite amount of negative energy
127: matter. The negative-mass \Schw has a qualitatively different
128: resolution interpreted to be fully reflecting. These 4d results
129: are similar to the 2d black hole and are generalized to an
130: arbitrary dimension $d>4$. }
131: 
132: %\keywords{}
133: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
134: \begin{document}
135: 
136: 
137: \section{Introduction}
138: 
139: The importance of singularities in General Relativity and their
140: resolution is well appreciated. The most intriguing ones are
141: considered to be those inside black holes, and the Big-Bang-like
142: cosmological singularities. String theory offers some cases where
143: a time-like singularity gets resolved: the orbifold, the flop, and
144: the conifold. These are all singularities in compact factors of
145: spacetime, the last two occurring in Calabi-Yau manifolds, and
146: their resolution often involves adding light matter (``twisted
147: sectors,'' ``wrapped D-branes'') which lives on the singularity
148: (for a review, see \cite{Polchinski} and references therein).
149:  Recently, there was a renewed effort to include
150: time-dependence in string theory (for a review, see
151: \cite{GRS-review} and references therein). Ultimately, one would
152: like to understand space-like singularities such as Schwarzschild
153: and cosmological singularities, but it would be fair to say that
154: there were no breakthroughs yet.
155: 
156: The study described here was motivated by bold attempts to cross
157: the singularity in a stringy black-hole model. The 2d black hole
158: is a case where an algebraic coset construction requires one to
159: add to spacetime additional manifolds which are glued together
160: over the singularity (see figures
161: \ref{PenroseRN},\ref{PenroseSchw}). For the uncharged 2d black
162: hole it was found \cite{dvv} that waves sent from the
163: ``additional'' infinity are fully reflected, while a charged 2d
164: black hole does have a non-zero transmission amplitude across the
165: singularity and, moreover, wave functions are smooth there
166: \cite{GiveonRabinoviciSever}.
167: 
168: Here we show that there is nothing intrinsically stringy in those
169: calculations~\footnote{There are $\alpha'$ corrections in string
170: theory, which were computed in the 2d case. These give rise to
171: extra phases in scattering amplitudes which vanish in the
172: semi-classical limit.}. Actually this interesting phenomenon
173: happens once one considers the wave equation on this background,
174: namely, by passing from point-particle probes~\footnote{The usual
175: definition of a singular spacetime is ``geodesic-incompleteness,''
176: namely, incomplete time-evolution for particle probes.} to waves.
177: More precisely, even though the wave equation is singular, the
178: ordinary differential equation (ODE) obtained after separating
179: time, and whose solutions we denote by $\phiom$, allows for a
180: unique, smooth continuation through the singularity.
181: Interestingly, this property is not special to the 2d charged
182: black hole, but it generalizes to the 4d charged
183: Reissner-Nordstr\"{o}m (RN) black hole, as well as to $d>4$ RN
184: black-holes (moreover, for the 2d and 4d cases the relevant ODE is
185: actually completely regular). This observation strongly suggests
186: that the above classical RN spacetimes should be viewed as being
187: made of two parts which are naturally connected across the
188: singularity at $r=0$. The two parts are
189:  (i) $r>0$ which is a usual (positive-mass) RN black
190: hole, and
191:  (ii) $r<0$, the region beyond the singularity,
192: which is a negative-mass spacetime.
193: 
194: Another motivation for considering the negative-$r$ part of RN
195: comes from the realistic astrophysical black holes, which are
196: known to be rotating in general \cite{bardeen,thorne,central}. In
197: the analytically-extended Kerr geometry (describing a stationary
198: spinning black hole), the Boyer-Lindquist coordinate $r$ goes
199: smoothly from positive to negative values. The region $r<0$ is an
200: asymptotically-flat universe (distinct from the original
201: positive-$r$ external universe). There is a curvature singularity
202: at $r=0$, but this singularity is a ring rather than a
203: hypersurface. A timelike geodesic heading towards $r=0$ will
204: generically avoid the ring singularity, and smoothly pass
205: ``through the ring'' into the negative-$r$ asymptotic region.
206: Obviously, wave packets will also pass through the ring, though
207: with partial reflection. Spherically-symmetric charged black holes
208: were often considered as useful toy models for the more realistic
209: spinning black holes, due to the remarkable similarity in the
210: inner structure of the two black-hole types. From this perspective
211: we may regard the RN solution analyzed here as such a simplified
212: toy model for a spinning black hole.
213: 
214: Initially one would expect the time evolution of a wave to be
215: ill-defined in the presence of a singularity since some boundary
216: condition (b.c.) needs to be supplied there. Given the smooth and,
217: in particular, univalued nature of the eigen-functions $\phiom(r)$
218: there is a natural way to formulate a unique time evolution simply
219: by following the usual recipe: decompose any incoming wave packet
220: into $\phiom$ and then endow each component with an $\exp(i\,
221: \omega\, t)$ time dependence. Since the eigen-functions
222: $\phiom(r)$ are smooth at $r=0$, it is just natural in this
223: formulation to consider both sides of the singularity to be
224: connected and communicating.
225: 
226: For negative-mass \Schw the situation is different from the
227: charged case, though our approach still provides a unique time
228: evolution. In this case the $r=0$ singularity ``flips'' from being
229: time-like at $r<0$ to space-like at $r>0$, suggesting that gluing
230: these two pieces would make little physical sense. Analyzing the
231: wave equation suggests that indeed in the Schwarzschild case the
232: field in the $r<0$ region evolves without any connection to the
233: $r>0$ region. Unlike the RN case, here the eigen-functions
234: $\phiom(r)$ generically have a log singularity, hence they cannot
235: be extended across the singularity in a univalued fashion. Yet, a
236: single regular $\phi_{1\omega}(r)$ solution exists for each
237: $\omega$. In this case a natural choice of boundary condition
238: suggests itself -- the so-called ``regularity b.c.'' -- i.e. the
239: demand that $\phi_{1\omega}(r)$ be bounded at the singularity.
240: Since we have a single regular solution for each $\omega$, any
241: wave packet coming from infinity will have a unique decomposition
242: into these regular functions $\phi_{1\omega}(r)$, and hence a
243: unique time evolution. (This is to be contrasted with the RN case,
244: in which there are two regular solutions for each $\omega$, and
245: therefore, the time evolution depends on initial data from both
246: sides of the singularity.) This ``regularity b.c.'' can be
247: physically interpreted as a reflecting b.c. While normally one can
248: define reflecting b.c. either by Dirichlet or by Neumann b.c. (or
249: a mix) the ``regularity b.c.'' at Schwarzschild's $r=0$ do not
250: leave us such a choice~\footnote{It was found already in
251: \cite{IshibashiHosoya} by different methods that there exists a
252: unique time evolution for negative-mass Schwarzschild, which is
253: probably the same one we explicitly describe and physically
254: interpret here.}.
255: 
256: Thus we find {\it a natural way to define (scalar wave) physics in
257: the presence of certain  time-like singularities} (by ``defining
258: physics'' we mean a recipe that allows to predict the time
259: evolution in spacetime) simply by passing from particle probes to
260: waves (we call such singularities ``wave-regular''). We discuss
261: two mechanisms: for RN we continue the space-time across the
262: singularity allowing for cross-communication, while for
263: negative-mass \Schw the ``regularity b.c.'' are unique and define
264: complete reflection.
265: 
266: 
267: We would like to pause to make a few comments on this picture:
268: \begin{itemize}
269: 
270: \item We consider this to be an important observation in the
271: search for a resolution of black hole singularities, at least
272: those of the time-like type. Yet, for this observation to be more
273: than a mere curiosity it is required that several additional tests
274: be satisfied such as regularity of higher-spin waves and
275: regularity against various perturbations to the equations (some of
276: which we perform).
277: 
278: 
279: \item At the classical level there are issues of interpretation,
280: for instance, the existence of these spacetimes or others with the
281: relevant singularities, the observers living in them and
282: implications to Cosmic Censorship, which we discuss in section
283: \ref{discussion}.
284: 
285: \item At the level of quantum gravity additional issues arise,
286: including  the validity of the wave equation and Hawking radiation
287: whose discussion we defer to section \ref{discussion} as well.
288: Here we only want to stress that one should be cautious not to
289: jump to conclusions and ``legitimize'' spacetimes with such
290: singularities as possible solitons.
291: 
292: \end{itemize}
293: 
294: After reviewing the RN and \Schw backgrounds in section
295: \ref{ReviewRN}, we explain our main observation in section
296: \ref{smooth-section}. We formulate the wave evolution for
297: arbitrary wave-packets coming from infinity (i.e. from large
298: negative $r$). However, the issue of specifying initial conditions
299: on a spacelike hypersurface crossing the $r=0$ singularity becomes
300: non-trivial and is relegated to section \ref{conditions-section}.
301: 
302: Once the unique evolution of the field is formulated, we are in a
303: position to investigate scattering phenomena. In section
304: \ref{cross-section-section} we study the physical properties of
305: the RN singularity by computing the cross section for
306: transmission. Even though the ODE for $\phiom$ is regular, the
307: effective potential diverges as $V_{\rm eff} \sim -c/r^{*~2}$ in
308: the vicinity of the singularity, where $r^*=r^*(r)$ is the
309: canonical ``tortoise'' coordinate. Therefore, one would expect the
310: singularity to have an effect even on  incoming waves of very high
311: energy -- unlike a regular point in spacetime, which becomes
312: ``transparent'' in this limit. Indeed we show that the high-energy
313: transmission amplitude for fixed angular momentum $l$ approaches
314: 25\% rather than 100\%. In this respect the singularity behaves as
315: a beam-splitter. Summing over $l$ we find the high-energy
316: dependence of the total cross section to be $\propto
317: \omega^{-4/3}$ where $\omega$ is the frequency of the incoming
318: plane wave. This behavior stands between the $\omega^0$ dependence
319: of absorption for a black hole with well-defined area, and the
320: $\omega^{-2}$ dependence of scattering off an elementary particle
321: in field theory.
322: 
323: Generalizations of the backgrounds to higher dimensions and for
324: the 2d black hole are presented in section \ref{generalizations}.
325: We determine the transmission amplitude and total cross-section as
326: in the 4d case. The only significant difference is that while in
327: 4d (and 2d) RN the variable-separated wave equation is regular,
328: for $d>4$ it is singular-regular. Yet, all the solutions
329: $\phiom(r)$ are still univalued for a continuation through the
330: singularity. For Schwarzschild, on the other hand, there are no
331: significant changes in passing from $d=4$ (and $d=2$) to $d>4$.
332: Hence, we also present the results for $d=4$ Schwarzschild in
333: section 5.
334: 
335: Next we challenge our picture by adding various perturbations in
336: section \ref{perturbations-section}. So far we discussed only
337: probes which were minimally-coupled scalar fields. Here we preform
338: several perturbations: first we consider a field with both mass
339: and charge, then we add interactions, and finally we consider the
340: effect of back-reaction.
341: 
342: 
343: In section \ref{conditions-section} we discuss some of the
344: conditions for a general timelike singularity to be wave-regular,
345: including the necessary constraint on initial conditions specified
346: on a spacelike hypersurface crossing the singularity, the
347: uniqueness of decomposition of an incoming wave-packet, and the
348: univalued property of the eigen-functions $\phiom$. In section
349: \ref{discussion} we summarize and discuss the issues mentioned
350: above, and list some open questions. Some technical details are
351: given in the appendices.
352: 
353: 
354: Note that, in principle, this study could have been carried out a
355: long time ago and, indeed, it turns out that there were some
356: attempts in this direction. However, it seems that none took the
357: conceptual freedom to cross the singularity and that actually led
358: several authors to pronounce RN to be wave-singular. Wald
359: \cite{Wald1980} gave a general discussion on extending the
360: definition of the wave equation to include non globally-hyperbolic
361: spacetimes relying on the mathematical notion of the ``Friedrichs
362: extension.'' Horowitz and Marolf \cite{HorowitzMarolf} showed that
363: some dilatonic black holes are wave-regular, where they consider a
364: Schr\"{o}dinger-like equation rather than the wave equation and
365: demonstrate ``essentially self-adjointness'' of the associated
366: operator. Ishibashi and Hosoya \cite{IshibashiHosoya} were able to
367: claim wave-regularity for negative-mass \Schw by going back to the
368: wave equation together with changing the metric in function space
369: accordingly from the ordinary $L^2$-metric to the Sobolev metric,
370: but using the same criterion of ``essentially self-adjointness.''
371: However, they do not comment on the physical nature of this
372: resolution. Related work also appears in \cite{Peeters:1994jz}.
373: Finally, Jacobson \cite{Jacobson} observed in passing the
374: regularity of the time-separated wave equation near the RN
375: singularity while studying the semiclassical decay of
376: near-extremal black holes.
377: 
378: {\bf Note added at second version:} it is interesting to consider
379: smooth approximations to the singular geometries. We found that if
380: one smoothly approximates the RN singularity one is required to
381: add a finite amount of negative energy matter, which is what one
382: would expect due to the wormhole nature of the singularity region.
383: The details of the calculation are given in section \ref{ReviewRN}
384: in the part titled ``Negative-energy content of the smeared
385: singularity.'' The subsequent ``perspectives on the $r<0$ region''
386: part in section \ref{ReviewRN} as well as the discussion section
387: should be read with this additional fact in mind. We gratefully
388: acknowledge the valuable comments of G. Gibbons  and J. Katz on
389: this matter. See \cite{Lynden-BellKatz,Lynden-BellKatz2} for a
390: closely related case.
391: 
392: \section{Review of the Reissner-Nordstr\"{o}m and Schwarzschild backgrounds}
393: \label{ReviewRN}
394: 
395: \vspace{0.5cm} \noindent {\it Basic features of the
396: Reissner-Nordstr\"{o}m geometry}
397: 
398: The background we wish to consider is the Reissner-Nordstr\"{o}m
399: (RN) geometry \cite{Reissner,Nordstrom}. In Schwarzschild
400: coordinates $(t,r,\theta ,\varphi )$ the metric is
401: \begin{equation}
402: ds^{2}=-f\,dt^{2}+f^{-1}\,dr^{2}+r^{2}\,d\Omega ^{2},
403: \label{RNmetric}
404: \end{equation}
405: where
406: \[
407: f=1-{\frac{2\,M}{r}}+{\frac{Q^{2}}{r^{2}}}
408: \]
409: and $d\Omega ^{2}$ is the two-sphere metric,
410: \[
411: d\Omega ^{2}=d\theta ^{2}+\sin ^{2}\theta \,d\varphi ^{2}.
412: \]
413: The electromagnetic four-potential is given by \footnote{The minus
414: sign was entered in order to conform with the rest of our sign
415: conventions and to produce (\ref{Veff-geod}) correctly.}
416: \[
417: A=-{\frac{Q}{r}}\,dt ~.
418: \]
419: For $Q=0$ the metric reduces to the Schwarzschild metric
420: \cite{Schw}.
421: 
422: For the (under-extremal) non-extremal black-hole case, $M>|Q|>0$
423: $f$ has two positive roots, located at
424: \[
425: r_{\pm }=M\pm \sqrt{M^{2}-Q^{2}}.
426: \]
427: Both $r=r_{+}$ and $r=r_{-}$ are null hypersurfaces, known as the
428: event horizon and the inner horizon, respectively (in
429: Schwarzschild there is only one such root, $r=r_{0}:=2\,M$, which
430: is the event horizon). Note that the roles of $r$ and $t$ as
431: space-like and time-like coordinates interchange at each root of
432: $f$.
433: 
434: The line element (\ref{RNmetric}) is singular at the horizons, but
435: this is merely a coordinate singularity which may be removed by
436: transforming to other coordinates (e.g. the Kruskal-Szekeres
437: coordinates discussed below). On the other hand, $r=0$ is a true
438: physical singularity (e.g. the Ricci scalar diverges in the RN
439: case), and we focus on this singularity in the present paper.
440: 
441: Usually one considers only the $r>0$ part of the RN geometry to be
442: physical, but {\it we will find it natural to consider the $r<0$
443: region as well}. Note a crucial difference in this regard between
444: the $r=0$ singularities in Reissner-Nordstr\"{o}m and in
445: Schwarzschild: in Reissner-Nordstr\"{o}m $f$ has a second-order
446: pole at $r=0$,
447: \begin{equation}
448: f\cong {\frac{Q^{2}}{r^{2}},}  \label{fRN-lead}
449: \end{equation}
450: and the $t$ coordinate is time-like on both sides of $r=0$, while
451: in Schwarzschild $f$ has a first-order pole
452: \begin{equation}
453: f\cong -{\frac{2\,M}{r}}  \label{fSchw-lead}
454: \end{equation}
455: and the coordinates ``flip'' their space-like/time-like nature:
456: $t$ is space-like for small and positive $r$ and it is time-like
457: for negative $r$. For this reason (as well as for other reasons),
458: we will later find it natural to glue the positive-$r$ and
459: negative-$r$ patches of Reissner-Nordstr\"{o}m over the
460: singularity, but not in the Schwarzschild case. The Penrose
461: diagram of the maximally extended Reissner-Nordstr\"{o}m
462: spacetime, including the $r<0$ region, is given in figure
463: \ref{PenroseRN}. This is to be contrasted with figure
464: \ref{PenroseSchw} which provides the Penrose diagrams of the
465: regions $r>0$ and $r<0$ in the Schwarzschild case. These diagrams
466: -- and particularly the fact that the singularity is spacelike at
467: $r>0$ and timelike at $r<0$ -- indicate the difficulty of matching
468: the two regions at $r=0$.
469: 
470: 
471: \vspace{.5cm} \noindent {\it Kruskal-Szekeres coordinates}
472: 
473: 
474: Although the Schwarzschild coordinates are perfectly suitable for
475: describing the $r<0$ region and its matching to $r>0$, insight
476: into the ``completed Penrose diagram'' may be achieved by using
477: the Kruskal-Szekeres coordinates, as we now describe. (In a first
478: reading one may skip the text and only view the completed Penrose
479: diagrams).
480: 
481: The $(r,t)$ coordinates (or ``\Schw coordinates'') define several
482: metric patches, which are separated on the $r$ axis by $r=0,\,
483: r_-,\, r_+$ in \RN and $r=0,\, r_0$ in Schwarzschild. In order to
484: see how these patches, and especially the $r<0$ patch, are
485: connected one transforms to the Kruskal-Szekeres coordinates
486: \cite{Kruskal,Szekeres}, and later performs a conformal
487: transformation to obtain the Penrose diagram (in appendix
488: \ref{KSapp} we review the procedure). Here again we will find a
489: qualitative difference between the \RN singularity and
490: Schwarzschild.
491: 
492: In \KS coordinates $(U,V)$, one uses a function $g(r)$ to
493: implicitly define $r$ through \be
494:  -U\, V = g(r)~. \ee
495: This function is defined by \be
496:  g(r) := \exp \(\pm 2\, \kappa\, r^*\)~, \ee
497: in terms of the ``tortoise'' coordinate $r^*$: \be
498:  dr^* := {dr \over f}~. \ee
499: $\kappa$ is the surface gravity of the relevant horizon and the
500: choice of sign is correlated with the choice of sign for $U$ and
501: $V$.
502: 
503: Let us look at the functions $g(r)$. For RN $g_\pm(r)$ are given
504: by (see figure \ref{gRNfigure}) \bea
505:  g_+(r) &=& \({r \over r_+}-1 \) \, \({r \over r_-}-1 \)^{-{\kappa_+ \over
506: \kappa_-}} \, \exp (2\, \kappa_+\, r) \non
507:  g_-(r) &=& \(1-{r \over r_-}\) \, \(1-{r \over r_+}\)^{-{\kappa_- \over
508: \kappa_+}} \, \exp
509:  (-2\, \kappa_-\, r) \eea
510: where $g_+$ covers the ``plus plane'' which includes the outer
511: horizon at $r_+: ~r_-<r<\infty$, and $g_-$ covers the ``minus
512: plane'' including the inner horizon at $r_-: ~-\infty<r<r_+$.
513: {}From the form of $g_-$ one realizes that in these coordinates
514: the $r<0$ region is naturally included and it lies at $(-U_-\,
515: V_-)
516: >1$, which is adjacent to the $0<r<r_-$ or $0 < -U_-\, V_- < 1$
517: region, and  as such we include it in the Penrose diagram (figure
518: \ref{PenroseRN}).
519: 
520: \begin{figure}
521: \centerline{\epsfxsize=70mm\epsfbox{gRN.eps}}
522: \medskip
523: \caption{The two functions $g_\pm(r)$ which determine $r$ in \KS
524: coordinates for \RN implicitly through $-U_\pm\, V_\pm = g_\pm
525: (r)$. $g_+$ is defined for the $r>r_-$, namely on the plus plane,
526: and $g_-$ is defined on $r<r_+$ namely on the minus plane. Note
527: that both functions are monotonic with range $(-\infty,+\infty)$
528: and hence there is always a unique solution for $r$. For $(-U_-\,
529: V_-)>1$ on the minus plane $r$ is negative. In drawing the figure
530: the values $r_-=.5, ~ r_+=1$ were used.}
531:  \label{gRNfigure}
532: \end{figure}
533: 
534: \begin{figure}
535: \centerline{\epsfxsize=90mm\epsfbox{PenroseRN.eps}}
536: \medskip
537: \caption{The complete Penrose diagram for the (under-extremal
538: $|Q|<|M|$) Reissner-Nordstr\"{o}m background including the regions
539: with $r<0$ which are not drawn usually. These regions are glued to
540: ``the other side'' of the time-like singularity, which is seen as
541: naked from that side. The wavy lines denote a typical scattering
542: experiment off the singularity. Waves from $\scri ^-$ are sent to
543: the singularity. Some of them are reflected back to $\scri ^+$
544: while the remainder is transmitted to $r>0$.} \label{PenroseRN}
545: \end{figure}
546: 
547: For Schwarzschild on the other hand, we have \be
548:  g_{\rm schw}(r) = \({r \over r_0}-1 \)\, \exp(r/r_0). \ee
549: One notes that unlike the $g_\pm$ which were monotonic $g_{\rm
550: schw}(r)$ has a minimum at $g_{\rm schw}(r=0)=-1$ (see figure
551: \ref{gSchw-figure}), exactly because $f_{\rm schw}(r)$ changes
552: sign at $0$. Thus for $-U\, V<-1$ there is no solution for $r$
553: while in the region $-1<-U\, V<0$ there are two solutions one of
554: them with negative $r$. Hence, the $r<0$ region is naturally
555: included and becomes a second cover over the inside of the black
556: hole ($0<r<r_0$) which we incorporate into the Penrose diagram in
557: figure \ref{PenroseSchw}. This second cover differs from \RN where
558: the $r<0$ region was located side by side with the ``ordinary''
559: $r>0$ regions. This difference is in tune with our later
560: conclusion that for \RN one should glue the two regions over the
561: $r=0$ singularity while in \Schw they should be considered to be
562: disconnected.
563: 
564: \begin{figure}
565: \centerline{\epsfxsize=70mm\epsfbox{gSchw.eps}}
566: \medskip
567: \caption{The function $g_{\rm schw}(r)$ which determines $r$ in
568: \KS coordinates (for the \Schw black hole) implicitly through
569: $-U\, V = g_{\rm schw} (r)$. Note that while for $g_{\rm schw}>0$
570: there is a unique solution for $r$, for $-1<g_{\rm schw}<0$ there
571: are two solutions, one of them with negative $r$ which lead to a
572: double cover in \KS coordinates. Finally for $g_{\rm schw}<-1$
573: there are no solutions.}
574:  \label{gSchw-figure}
575: \end{figure}
576: 
577: 
578: \begin{figure}
579: \centerline{\epsfxsize=120mm\epsfbox{PenroseSchw.eps}}
580: \medskip
581: \caption{The complete Penrose diagram for a Schwarzschild black
582: hole including the negative-mass regions which appear as a double
583: cover of the inside of the black (or white) hole. The
584: negative-mass regions are shown separately on the side, rotated by
585: 90 degrees so that time flows in them upward as is conventional.
586: Here we will find no communication through the singularity.}
587: \label{PenroseSchw}
588: \end{figure}
589: 
590: \vspace{0.5cm}
591:  \noindent {\it Negative-energy content of the smeared singularity}
592: 
593: The electro-vacuum geometry constructed here has a curvature
594: singularity at $r=0$. One may consider modifying the metric
595: function in the very neighborhood of $r=0$, say at some $|r|\leq
596: r_{0}$, so as to make the geometry smooth. In this way one obtains
597: a wormhole smoothly connecting the positive-$r$ and the
598: negative-$r$ parts. However, any such smearing will imply the
599: presence of additional energy-momentum in the region $|r|\leq
600: r_{0} $, as would be determined by applying the Einstein operator
601: to the modified metric. The specific form of the additional
602: energy-momentum tensor will depend on the specific choice of
603: smearing. It is well known, however, that any wormhole of this
604: kind requires ``exotic'' matter fields, i.e. an energy-momentum
605: source which violates the weak energy condition. A simple way to
606: understand it is to consider a congruence of null rays propagating
607: inward radially. In this setting Raychaudhuri's equation tells us
608: that as long as the energy condition is satisfied ${d^2 \over
609: d\lambda^2} \log(A) \le 0$, where $A$ is the transverse area (of
610: the two-sphere) and $\lambda$ is an affine parameter along the
611: ray. Since $A$ (and $\log(A)$) is asymptotically decreasing in an
612: inward radial motion, but after going through the ``wormhole'' it
613: is again increasing, there must be an intermediate region where
614: the weak energy condition is violated.
615: 
616: We wish to evaluate the amount of negative energy required to
617: support this wormhole-shaped smeared geometry, particularly at the
618: limit $r_{0}\to 0$. To be more specific, we consider here a
619: thin-shell model \cite{Israel}. Namely, we directly match the
620: geometry at $r\geq r_{0}$ to that at $r\leq -r_{0}$, through a
621: thin shell located at $|r|=r_{0}$, for some $0<r_{0}\ll Q$. This
622: thin shell is a timelike hypersurface parameterized by the three
623: coordinates $(t,\theta ,\varphi )$. Note that the induced
624: three-geometry is continuous across this shell: $g_{\theta \theta
625: }$ and $g_{\varphi \varphi }$ only depend on $r^{2}$, and $g_{tt}$
626: (which originally has different values at the two sides) becomes
627: continuous after a trivial rescaling of the $t$ coordinate at e.g.
628: $r<-r_{0}$.
629: 
630: The shell's contribution to the energy-momentum tensor may be
631: expressed as \cite{Visser}
632: \begin{equation}
633: T^{\mu \nu }(x)=S^{\mu \nu }(x)\delta (\eta ) ~,
634: \end{equation}
635: where $S^{\mu \nu }$ is the shell's surface energy-momentum
636: distribution, and $\eta $ is the proper-length parameter along
637: radial lines of constant $t,\theta ,\varphi $, with $\eta =0$ at
638: the shell and $\eta >0$ at $r>r_{0}$. The surface distribution
639: $S^{\mu \nu }$ is determined from the jump in the extrinsic
640: curvature,
641: \begin{equation}
642: S_{j}^{i}=-\frac{1}{8\pi }\left( \left\langle K\right\rangle
643: _{j}^{i}-\delta _{j}^{i}\left\langle K\right\rangle
644: _{k}^{k}\right) .
645: \end{equation}
646: Hereafter $i,j,k$ run over the three coordinates $(t,\theta
647: ,\varphi )$, and
648: \begin{equation}
649: \left\langle K\right\rangle _{j}^{i}\equiv
650: K_{j}^{i(+)}-K_{j}^{i(-)},
651: \end{equation}
652: where $K_{j}^{i(\pm )}$ is the shell's extrinsic curvature with
653: respect to the geometries at $r>r_{0}$ and $r<-r_{0}$,
654: respectively. (We are using here General-Relativistic units,
655: $G=c=1$.) The extrinsic curvature may conveniently be expressed in
656: terms of Israel's ``natural coordinates'' \cite{Israel}, which we
657: take here to be the three hypersurface coordinates $t,\theta
658: ,\varphi$ and the above proper-length coordinate $\eta $:
659: \begin{equation}
660: K_{ij}^{(\pm )}=\frac{1}{2}\left[ \frac{\partial g_{ij}}{\partial
661: \eta } \right] _{(\pm )}=\frac{1}{2}\left[
662: g_{rr}^{-1/2}\frac{\partial g_{ij}}{
663: \partial r}\right] _{(\pm )}  ~, \label{extrinsic}
664: \end{equation}
665: with the derivatives (and $g_{rr}$) evaluated at the $(+)$ or
666: $(-)$ sides of the shell, respectively.
667: 
668: We are primarily interested here in the amount of (negative)
669: energy contributed by the shell, represented by $S_{t}^{t}$:
670: \begin{equation}
671: S_{t}^{t}=\frac{1}{8\pi }\left( \left\langle K\right\rangle
672: _{\theta }^{\theta }+\left\langle K\right\rangle _{\varphi
673: }^{\varphi }\right) =\frac{ 1}{4\pi }\left\langle K\right\rangle
674: _{\theta }^{\theta } ~.
675: \end{equation}
676: A straightforward application of Eq. (\ref{extrinsic}) yields
677: \begin{equation}
678: K_{\theta }^{\theta (\pm )}=\left[
679: \frac{\sqrt{1-2M/r+Q^{2}/r^{2}}}{r} \right] _{r=\pm r_{0}} ~,
680: \end{equation}
681: hence
682: \begin{equation}
683: \left\langle K\right\rangle _{\theta }^{\theta }=\frac{\sqrt{
684: 1-2M/r_{0}+Q^{2}/r_{0}^{2}}+\sqrt{1+2M/r_{0}+Q^{2}/r_{0}^{2}}}{r_{0}}\,.
685: \end{equation}
686: 
687: Let $\rho $ denote the shell's energy density as measured by a
688: static observer:
689: \begin{equation}
690: \rho =-T_{t}^{t}=-\frac{1}{4\pi }\left\langle K\right\rangle
691: _{\theta }^{\theta }\delta (\eta )
692: \end{equation}
693: (recall that $T_{t}^{t}$ is invariant to a rescaling of $t$). The
694: total shell's energy is obtained by integrating over the 3-volume,
695: i.e. over $\eta $ and over the shell's area:
696: \begin{eqnarray*}
697: E_{shell} &=&4\pi r_{0}^{2}\int \rho d\eta =-r_{0}^{2}\left\langle
698: K\right\rangle _{\theta }^{\theta } \\
699: &=&-r_{0}\left( \sqrt{1-2M/r_{0}+Q^{2}/r_{0}^{2}}+\sqrt{
700: 1+2M/r_{0}+Q^{2}/r_{0}^{2}}\right) .
701: \end{eqnarray*}
702: 
703: Finally, we calculate $E_{shell}$ at the limit where $r_{0}$
704: shrinks to zero:
705: \begin{equation}
706: E_{shell}^{0} \equiv \lim_{r_0 \to 0} E_{shell}=-2Q.
707: \end{equation}
708: 
709: %\newpage
710: \vspace{0.5cm}
711:  \noindent {\it Perspectives on the $r<0$ region}
712: 
713: The metric (\ref{RNmetric}) is invariant under the transformation
714: $r\to -r,~M\to -M,$ and hence we shall often refer to the $r<0$
715: region as the ``negative mass'' region. To an observer in the
716: asymptotically-flat region at $r<0$ the central object will appear
717: as one of negative mass, namely, a gravitationally-repelling
718: object.\footnote{This is also the situation in the Kerr geometry:
719: in the asymptotically-flat region $r<0$ the central object is
720: gravitationally repelling.}
721: 
722: We alert the reader that negative masses have very unusual
723: features:
724: 
725: \begin{itemize}
726: \item  A negative mass produces anti-gravity, namely it repels all
727: masses.
728: 
729: \item  As a result of Newton's second law $F=m\, a$, the
730: acceleration of a negative mass is reversed to the force acting on
731: it.
732: 
733: \item  As an immediate result of the two properties above, if one
734: were to place two bodies initially at rest, one with a negative
735: mass and the other with a positive mass, both will accelerate in
736: the same direction going from the negative mass to the positive
737: one (furthermore, if the two masses are of the same magnitude,
738: they will uniformly accelerate forever).
739: 
740: \end{itemize}
741: 
742: 
743: Due to these unusual features (and possibly others), negative
744: masses are often excluded (however, see
745: \cite{negative-supergravity}). We shall not attempt here to
746: provide a conclusive answer to this question. Instead, we will
747: take a pragmatic approach of exploration by allowing negative-mass
748: spaces in the current context and being alert to the appearance of
749: a problem such as a resulting inconsistency. We will not find any
750: such inconsistency in this paper, and so we regard this question
751: as remaining open.
752: 
753: 
754: In the (under extremal) RN black-hole case, $M>|Q|>0$, one can
755: reach the $r=0$ singularity either from the negative-mass
756: asymptotic region (i.e. from large negative $r$), or from $r \to
757: r_{-}$, the inner horizon. Looking at the Penrose diagram of the
758: maximally extended spacetime, figure \ref{PenroseRN}, one may
759: consider reaching the inner horizon by jumping into the black hole
760: from the ordinary (positive-mass) asymptotic region. The inner
761: horizon of the pure RN geometry is a perfectly-regular
762: hypersurface. However, it is known already for some time that the
763: inner horizon is unstable to perturbation and should become
764: singular \cite{Penrose,PI} in ``realistic'' charged black holes.
765: For this reason it was often believed that the inner horizon
766: cannot be crossed. Recent investigations, however, indicated that
767: the singularity at the inner horizon is weak, that is, the tidal
768: forces are too weak to harm physical objects
769: \cite{weak-singularity,burko,spinning}. Therefore, there is no
770: obvious reason to exclude objects falling into the black hole and
771: arriving at $r=0$ through the inner horizon. For conceptual
772: simplicity, however, we shall primarily consider here observers
773: arriving at $r=0$ from the negative-$r$ asymptotically-flat
774: region, namely, from beyond the singularity.
775: 
776: In the case $|Q|>M$ (figure \ref{PenroseRN-ext}) there are no
777: horizons, but simply two asymptotically-flat regions, one with
778: positive mass and the other with negative mass (in this case there
779: is no reason to transform into Kruskal-Szekeres -like
780: coordinates). In both regions there is a naked singularity at
781: $r=0$, and we glue the two patches there.
782: 
783: \begin{figure}
784: \centerline{\epsfxsize=55mm\epsfbox{PenroseRN-ext.eps}}
785: \medskip
786: \caption{The complete Penrose diagram for an over-extremal
787: ($|M|<|Q|$) Reissner-Nordstr\"{o}m black hole. It is composed of a
788: positive-mass RN glued over the singularity to a negative-mass RN.
789: There are no horizons and the singularity is naked on both sides.}
790: \label{PenroseRN-ext}
791: \end{figure}
792: 
793: \vspace{0.5cm} \noindent {\it Geodesics}
794: 
795: We end this section with a review of particle motion in this
796: background, and a derivation of the effective potential for
797: particles (or geodesics) which we later compare with the effective
798: potential for waves. The geodesic equation for a particle of mass
799: $m$ can be derived from the Hamiltonian system
800: \begin{equation}
801: {\cal H}={\frac{1 }{2}}\, g^{\mu\, \nu}\, p_\mu\, p_\nu =
802: -{\frac{1 }{2}} \, m^2 ~~,
803: \end{equation}
804: where the Hamiltonian is automatically with respect to a multiple
805: of the proper time $\tau$. For a particle of charge $q$ one
806: replaces $p \to p-q\, A$.
807: % sign as in Landau-Lifshitz QM (113.1)
808: 
809: Since ${\cal H}$ is independent of $t$ and $\,\phi $ we have the
810: constants of the motion $E:=-p_{t},\,l:=p_{\phi }$. The motion is
811: then described by
812: \begin{eqnarray}
813: 0 &=&\dot{r}^{2}+V_{{\rm eff}}^{{\rm geodesic}}(r)~,  \nonumber \\
814: V_{{\rm eff}}^{{\rm geodesic}}(r) &=&\left({\frac{l^{2}}{r^{2}}}\,
815: +m^{2}\right)\,f(r)-\left(E-{\frac{q\,Q}{r}}\right)^{2}~.
816: \label{Veff-geod}
817: \end{eqnarray}
818: In particular, around $r\simeq 0$, $V_{{\rm eff}}\sim
819: {\frac{l^{2}Q^{2}}{r^{4}}}$ for $l>0$ [and $V_{{\rm eff}}\sim
820: (m^{2}-q^{2}){\frac{Q^{2}}{r^{2}}}$ for $l=0$]. Note that the
821: effective potential given here does not have the ordinary units of
822: potential, but it has the advantage of being general and valid
823: both for massive and massless particles. In order to get the
824: physical potential in a Newtonian approximation one needs to
825: rescale $V_{{\rm eff}}$ by an appropriate power of $m$.
826: 
827: 
828: 
829: \section{Waves are smooth}
830: \label{smooth-section}
831: 
832: \subsection{Set-up}
833: 
834: A spacetime is usually defined to be singular when it is not
835: geodesically complete, namely with respect to point-particle
836: probes. Yet, all the fundamental interactions are formulated for
837: field theories, where particles are derived concepts constructed
838: from appropriate wave packets. Moreover, the mashing of Quantum
839: Mechanics and Special Relativity requires the quantum theory of
840: fields. Thus it is natural to probe spacetimes with waves rather
841: than point-particles.
842: 
843: Here we will confine ourselves, as a first step, to scalar fields.
844: Initially we will further assume a minimally-coupled free field
845: with no mass or charge, and later these assumptions will be
846: relaxed and shown not to affect the results.
847: 
848: When should we consider a spacetime to be wave-regular? A central
849: objective of physics is to provide predictions for the results of
850: experiments, or a well-defined time evolution, and it is exactly
851: this predictability which is at risk in the presence of a
852: singularity. We would therefore consider a spacetime to be
853: wave-regular if the time evolution is unique for any physically
854: acceptable initial conditions (to be discussed below). Technically
855: the issue is whether one can supply boundary conditions at the
856: singularity such that the time evolution is regular and whether
857: these are unique.
858: 
859: In ``normal'' (i.e. non-singular) situations there are two
860: different set-ups for initial conditions: (i) the Cauchy
861: formulation, in which the field and its time derivative are
862: specified on a spacelike hypersurface; and (ii) the characteristic
863: formulation, in which the field is specified on two null
864: hypersurfaces. This second formulation has a special, widely used
865: variant (iia) in which the null hypersurfaces are taken to be at
866: the null past boundaries of the region of spacetime under
867: consideration -- e.g. past null infinity (for an
868: asymptotically-flat region), and a past horizon (for an
869: asymptotically-flat region outside a black hole).
870: 
871: In our problem -- formulating the time evolution of the field at
872: the two sides of the $r=0$ singularity -- the Cauchy formulation
873: (i) has a potential problem, because any spacelike initial
874: hypersurface must cross the $r=0$ singularity. It is a priori
875: unclear what are exactly the ``physically-acceptable initial
876: conditions'' near $r=0$. On the other hand, the characteristic
877: formulation (iia) is free of this potential problem. The two null
878: hypersurfaces which naturally suggest themselves for this
879: formulation are past null infinity ($\scri^-$) of the negative-$r$
880: universe, and the inner horizon $r_{-}$. These two hypersurfaces
881: nowhere intersect the singularity, or even get close to it.
882: 
883: We shall therefore base our analysis of time evolution on this
884: characteristic formulation. That is, we probe the spacetime by
885: wave-packets prepared and sent in from the asymptotic regions far
886: away from the singularity. Indeed, it is physically reasonable to
887: characterize the singularity by its response to any possible
888: scattering experiment. Motivated by these considerations, we shall
889: define a spacetime (or a region of spacetime) to be {\it
890: wave-regular} if the time evolution is unique for generic initial
891: conditions specified at the relevant past null boundaries of this
892: (region of) spacetime.
893: 
894: While the characteristic formulation (ii) avoids the issue of
895: putting constraints on initial conditions near the singularity,
896: which is indeed non-trivial in hindsight, it does require a global
897: point of view, whereas our resolution is essentially local. This
898: local property is better understood in the Cauchy formulation (i)
899: and hence we will briefly review both formulations in this
900: section. Actually one could define also a hybrid of these methods
901: which keeps both locality and some isolation from the singularity:
902: namely, a formulation (ia) in which the field and its derivatives
903: are supplied on a spacelike hypersurface but are required to
904: vanish in a neighborhood of the singularity.~\footnote{This
905: variant is slightly more restrictive than just a re-formulation of
906: the initial-value setup, as it demands that the field strictly
907: vanishes at some open neighborhood. Nevertheless it is still
908: sufficiently general to allow for physically-meaningful scattering
909: experiments.}
910: 
911: In what follows we shall show that both the $-\infty < r < r_-$
912: piece of the RN spacetime and negative-mass \Schw are indeed
913: wave-regular. We shall define first the time evolution in a
914: straightforward manner in the Cauchy formulation, and then after
915: defining a canonical form for the radial equation and the
916: effective potential we shall state the characteristic formulation.
917: A treatment of the interesting subtleties in the Cauchy
918: formulation will be deferred to section \ref{conditions-section}.
919: 
920: \subsection{Separation of variables}
921: 
922: Let us proceed by writing down the wave equation for a minimally
923: coupled massless scalar field $\phi $ in the RN (or Schwarzschild)
924: background (\ref {RNmetric}). After separating the angular
925: variables \be
926:  \phi(r,t,\Omega) = \phi_l(r,t)\, Y_{lm}(\Omega)~, \ee
927:  where $\Omega$ denotes the angular variables~\footnote{The notation
928: $\Omega$ for the angular variables is convenient for the
929: generalization to higher dimensions where there are more angular
930: variables and there is no need to explicitly define them.}
931:  $(\theta,\, \varphi)$, one has
932: \begin{equation}
933: 0=\Box \phi =\left[ -f^{-1}\,{\partial
934: }_{tt}+{\frac{1}{r^{2}}}\,{\partial }_{r}\,f\,r^{2}\,{\partial
935: }_{r}-{\frac{l\,(l+1)}{r^{2}}}\right] \,\phi _{l}~.
936: \label{waveeq1}
937: \end{equation}
938: The wave equation (\ref{waveeq1}) is singular at $r=0$: after
939: multiplying the equation by $f$ to extract ${\partial }_{tt}$, the
940: part containing e.g. the second-order $r$ derivative becomes
941: $f^{2}\,{\partial }_{rr}$, and $f^{2}$ diverges as $r^{-4}$. This
942: is not surprising since we know that $r=0$ is a curvature
943: singularity.
944: 
945: Still, we may attempt a straight-forward separation of the time
946: variable,
947: \begin{equation}
948: \phi _{l}(r,t)=\phi _{\omega l}(r)\,{\rm exp}(i\,\omega \,t)~,
949: \label{basicseparation}
950: \end{equation}
951: getting \bea
952:  0 &=& \Box_{\omega l}\, \phi_{\omega l} \non
953:  \Box_{\omega l} &:=& f^{-1}\, \omega^2+
954:  {\frac{1}{r^{2}}}\,{\partial }_{r}\,f\,r^{2}\,{\partial }_{r}-
955:  {\frac{l\,(l+1)}{r^{2}}}~. \eea
956: It will be convenient to normalize $\Box_{\omega l}$ in two
957: different ways: if we want to have a unit coefficient for
958: ${\partial }_{rr}$ then we work with $f^{-1}\, \Box_{\omega l}$
959: while $f\, \Box_{\omega l}$ achieves a unit coefficient for
960: $\omega^2$. Indeed, writing down $f^{-1}\, \Box_{\omega l}$
961: explicitly, we have \bea
962:  0 &=& f^{-1}\, \Box_{\omega l}\, \phi_{\omega l} =\non
963:  &=& \left[ {\frac{1}{f\,r^{2}}}\, \partial_{r}\,f\, r^2\, \partial
964: _r-{\frac{l\,(l+1)}{f\,r^2}} + f^{-2}\, \omega ^{2}\right]
965: \,\phi_{\omega l} ~~, \label{separation-of-var} \eea and at this
966: point we need to discuss the Reissner-Nordstr\"{o}m case and the
967: Schwarzschild case separately.
968: 
969: \vspace{.5cm} \noindent {\it Reissner-Nordstr\"{o}m }
970: 
971: Since $f_{RN}\,r^{2}\cong Q^{2}$ is regular at $r=0$, the
972: variable-separated equation (\ref{separation-of-var}) is
973: surprisingly completely regular. This observation is at the heart
974: of our proposed resolution. To leading order it is given by
975: \begin{equation}
976: 0=\left[ (1+\dots )\,{\partial }_{rr}+O(r^{0})\,{\partial
977: }_{r}-(1+\dots )\,{\frac{l\,(l+1)}{Q^{2}}}+(1+\dots
978: )\,{\frac{r^{4}}{Q^{4}}}\,\omega ^{2}\right] \phi _{\omega}~,
979: \label{rnreg}
980: \end{equation}
981: where ellipses always denote corrections which are higher order in
982: $r$, and from now we will often write $\phiom$ instead of
983: $\phi_{\omega l}$ suppressing the index $l$ for clarity. Note that
984: the equation is not only non-singular but also analytic (namely,
985: $\phi _{\omega ,rr}$ is an analytic function of $\phi _{\omega
986: },\phi _{\omega ,r}$ and $r$) at $r=0$, for any $l$ and $\omega $.
987: 
988: Since all the solutions for this equation are smooth at $r=0$, a
989: natural way to define the time evolution suggests itself: given an
990: initial wave packet (which propagates from either $\scri^-$ and/or
991:  the inner horizon) decompose it into radial eigen-functions $\phi
992: _{\omega }(r)\,$. Since the eigen-functions $\phi _{\omega }(r)$
993: are smooth at $r=0$, the time evolution defined above will be
994: smooth there. Consequently, parts of the wave packet will cross
995: this point. It is therefore essential to include both sides of the
996: $r=0$ singularity in our analysis.
997: 
998: In equations: given initial conditions $\phi_i(r):=\phi(r,t=0)$
999: and $\dot{\phi}_i(r):=\dot{\phi}(r,t=0)$ (we chose here the
1000: initial hypersurface $t=0$ without loss of generality), one
1001: determines the decomposition $\phi_+(k),\, \phi_-(k)$ that
1002: satisfies \bea
1003:  \phi_i(r) &=& \int \( \phi_+(k) + \phi_-(k)\, \)\, \phi_k(r)\, dk \non
1004:  \dot{\phi}_i(r) &=& \int i \omega\, \( \phi_+(k) - \phi_-(k)\, \)\,
1005: \phi_k(r)\, dk
1006:  \eea
1007: where here $\omega:=|k|$, and $\phi_{k}$ denotes two independent solutions
1008: (one for positive $k$ and one for negative $k$)
1009: of the wave equation $\Box_{\omega l}=0$ (with $\omega=|k|$).
1010:  \footnote{$\phi_{k}$
1011: are usually chosen such that $\phi_k \simeq \exp (i\, k\,
1012: r^*\,)/r$ at large $r^*$ (large negative $r$).}
1013:  Now the time evolution is defined naturally to be \footnote{This amounts to
1014:  taking the radial function $\phi _{\omega l}(r)$
1015:  of Eq. (\ref{basicseparation}) to be
1016:  $\phi_+(k=+\omega)+\phi_+(k=-\omega)$ for $\omega>0$ and
1017:  $\phi_-(k=+\omega)+\phi_-(k=-\omega)$ for $\omega<0$.}
1018:  \be
1019:  \phi(r,t) = \int \( \phi_+(k)\,  \exp(+i\, \omega\, t)
1020:  + \phi_-(k)\,  \exp(-i\, \omega\, t)\, \) \phi_k(r)\, dk ~~.\ee
1021: 
1022: In the characteristic formulation the decomposition into modes may
1023: be done by Fourier-decomposing the waves coming from $\scri^-$ or
1024: from the inner horizon in ${\rm exp}(i\,\omega \,v)$ or ${\rm
1025: exp}(i\,\omega \,u)$, respectively, where $u$ and $v$ are the two
1026: Eddington-like null coordinates, and associating a certain radial
1027: function $\phi _{\omega }(r)$ with each Fourier component. This
1028: will be described in more detail in subsection
1029: \ref{char-subsection}.
1030: 
1031: 
1032: At first it may look as if the singular partial differential
1033: equation (\ref {waveeq1}) has turned completely regular. Later we
1034: will see, however, that some imprint of the singularity remains.
1035: This is expressed both in the unusual large-$\omega$ limit of the
1036: transmission amplitude through $r=0$ (see section
1037: \ref{cross-section-section}) and in the constraints for Cauchy
1038: initial conditions at $r \sim 0$ (see section
1039: \ref{conditions-section}).
1040: 
1041: \vspace{.5cm} \noindent {\it Schwarzschild }
1042: 
1043: Using the leading behavior of $f_{{\rm Schw}}$ at $r=0$
1044:  (\ref{fSchw-lead}), the variable-separated equation
1045: (\ref{separation-of-var}) is given to leading order by
1046: \begin{equation}
1047: 0=\left[ (1+\dots )\,{\partial }_{rr}+{\frac{(1+\dots
1048: )}{r}}\,{\partial } _{r}+(1+\dots
1049: )\,{\frac{l\,(l+1)\,}{2\,M\,r}}+(1+\dots
1050: )\,{\frac{r^{2}}{4\,M^{2}}}\,\omega ^{2}\right] \phi _{\omega } ~.
1051: \label{Schw-var-sep}
1052: \end{equation}
1053: One notices that this equation is not regular anymore, but rather
1054: the $\partial_r$ and the $l(l+1)$ terms are singular. Actually it
1055: falls into the so called ``regular-singular'' category of mild
1056: singularities (see appendix \ref{reg-sing-app} for a rudimentary
1057: review).
1058: 
1059: Looking at Eq. (\ref{Schw-var-sep}) one realizes that the
1060: characteristic exponents (defined by the attempted leading-order
1061: solution $\phi _\omega \cong r^{\rho }$) are $\rho
1062: _{1}=\rho_{2}=0$, and thus the two independent solutions behave as
1063: $\phi _{1 \omega}\sim r^{0},~\phi _{2 \omega}\sim \log (r)$ at
1064: $r\sim 0$. So unlike Reissner-Nordstr\"{o}m one of the solutions
1065: is singular, and moreover does not have a single-valued
1066: continuation to negative $r$, and there is no sense in gluing
1067: together positive $r$ to negative $r$. It is perhaps fortunate
1068: that this is the case, since gluing the two sides would require
1069: dealing with the ``flip'' in the nature of $(r,t)$ from space-like
1070: to time-like (and an arbitrary choice of the direction of the
1071: arrow of time). Still we find it remarkable that one of the
1072: solutions does turn out to be regular. This half-regularity allows
1073: a natural definition of the time evolution, as we shortly
1074: describe, where this time there is no transmission through $r=0$.
1075: 
1076: The situation here is analogous to the relation between a free
1077: field $\phi $ on the infinite line $-\infty <r<+\infty $ to a free
1078: field on the semi-infinite line $0<r<+\infty $, which could be
1079: e.g. the radial coordinate in some higher dimension, with
1080: Dirichlet or Neumann boundary condition at $r=0$. For the infinite
1081: line there are two $\phi_{\omega }$ solutions for each $\omega$,
1082: which allow a decomposition of wave packets arriving
1083: simultaneously from both $-\infty$ and $+\infty$. On the other
1084: hand, for the semi-infinite line the boundary condition selects a
1085: single $\phi _{\omega }$ solution for each $\omega $. This still
1086: allows for a decomposition of a wave packet, which this time can
1087: arrive only from $+ \infty $. The situation for
1088: Reissner-Nordstr\"{o}m is similar to the infinite line, while
1089: negative-mass Schwarzschild is similar to the semi-infinite line
1090: (we consider the $r<0$ side of the Schwarzschild singularity since
1091: there $t$ is time-like and one can consider scattering
1092: experiments).
1093: 
1094: Thus we naturally define the time evolution for negative-mass
1095: Schwarzschild by decomposing the initial conditions (defined only
1096: at $r=-\infty $) into the regular eigen-functions $\phi _{1
1097: \omega}$ alone and then proceeding as usual to define the time
1098: evolution. Since $\phi _{1 \omega}$ is smooth at $r=0$, the time
1099: evolution will be smooth there too. As it turns out it was found
1100: already in \cite{IshibashiHosoya} by somewhat formal arguments
1101: that there exists a unique time evolution for negative-mass
1102: Schwarzschild, which is probably the same as the one we just
1103: explicitly described. For RN on the other hand we differ as we get
1104: wave-regularity after considering both sides of the singularity,
1105: while they do not.
1106: 
1107: \sbsection{Perspectives on the boundary conditions for \Schw}
1108: 
1109: In the \Schw case we simply impose the boundary condition that the
1110: scalar field be regular at $r=0$.
1111: This does not conform with either the Dirichlet or Neumann b.c.;
1112: namely, neither $\phi$ nor its radial derivative vanishes at $r=0$.
1113: In fact, the Dirichlet or Neumann b.c. are both inconsistent with
1114: the asymptotic behavior near the singularity (due to the presence of a
1115: divergent mode).
1116: 
1117: Yet, there is a physically appealing procedure
1118: to obtain our regularity b.c.
1119: from the Dirichlet or Neumann b.c. (or a mixture).
1120: We can cutoff space at some $r_1<0$ close to the
1121: singularity, and then take the limit $r_1 \to 0$. Since all modes are
1122: well-behaved at $r=r_1$,
1123: we are free to impose either Dirichlet or Neumann b.c. there.
1124: Let us denote as before the
1125: solution regular at $r=0$ by $\phi_{1 \omega}$, and some other independent and
1126: necessarily singular solution by $\phi_{2 \omega}$. We are looking
1127: for a linear combination of the two, $\phiom = a\, \phi_{1 \omega}
1128: + b\, \phi_{2 \omega}$, that will satisfy the boundary conditions at
1129: $r_1$. If we impose Dirichlet we need \be
1130:  0=\phiom(r_1) = a\, \phi_{1 \omega}(r_1) +b\,
1131:  \phi_{2 \omega}(r_1) ~. \ee
1132:  As we take $r_1 \to 0$ $\phi_{1 \omega}$
1133:  remains finite while $\phi_{2 \omega}$
1134: diverges with a log singularity, and hence we must take $b=0$,
1135: namely the ``regular solution b.c.''
1136: 
1137: If we choose Neumann conditions instead, we note that again while
1138: $\phi_{1 \omega}'$ remains finite $\phi_{2 \omega}'$ diverges with
1139: a $1/r$ singularity and hence we are again forced to take $b=0$.
1140: Altogether we find that in both cases as $r_1 \to 0$ we are left
1141: with a unique b.c., namely our ``regular b.c.''
1142: 
1143: \subsection{Normal form of the radial equation}
1144: 
1145: \label{normal-form-subsection}
1146: 
1147: Transforming to a new radial coordinate (the so called ``tortoise
1148: coordinate'') $r^*(r)$ defined by
1149: \begin{equation}
1150: dr^{*}={\frac{dr}{f(r)},} \label{def-rstar}
1151: \end{equation}
1152: and to a new field variable
1153: \begin{equation}
1154: \phi ^{*}:= r\phi  \label{defn-g}
1155: \end{equation}
1156: (and correspondingly $\phi _{\omega }^{*}:= r\phi _{\omega }$),
1157: the wave equation (\ref{waveeq1}) and the radial equation
1158: (\ref{separation-of-var}) take the standard forms
1159: \begin{equation}
1160: f^{-1}\, \left[ {\partial }_{tt}-{\partial }_{r*r*}+V_{{\rm
1161: eff}}(r^*)\right] \phi^*=0 \label{normal-full}
1162: \end{equation}
1163: and
1164: \begin{equation}
1165: \left[ -{\partial }_{r*r*}+V_{{\rm eff}}(r^*)-\omega ^{2}\right]
1166: \phi _{\omega }^{*}=0.  \label{normal-radial}
1167: \end{equation}
1168: Here the effective potential $V_{{\rm eff}}$ is given by $V_{{\rm
1169: eff}}^{\rm geodesic}$, the effective potential for particles,
1170: with the substitutions $l^2 \to l(l+1)$ and $E \to 0$,
1171: \footnote{in principle $E$ is to be now replaced by $\omega$, but
1172: the latter already appears explicitly in Eq.
1173: (\ref{normal-radial}). Recall also that we assume at this stage a
1174: neutral massless scalar field, hence $m=q=0$.} plus an extra term
1175: $\Delta V_{{\rm eff}}$:
1176: \begin{equation}
1177: V_{{\rm eff}}(r^*)=\Delta V_{{\rm eff}}+f\,{\frac{l(l+1)}{r^{2}},}
1178: \label{Veff}
1179: \end{equation}
1180: where
1181: \begin{equation}
1182: \Delta V_{{\rm eff}}(r^*)={\frac{f}{r}}\,{\partial }_{r}\,f ~.
1183: \label{DeltaVeff}
1184: \end{equation}
1185: 
1186: We find it instructive to explore the leading terms in the radial
1187: equation at $r=0$, when this equation is expressed in terms of
1188: $r^*$ rather than $r$.
1189: 
1190: We start by looking at the Reissner-Nordstr\"{o}m case. To leading
1191: order near $r\sim 0$
1192: \begin{equation}
1193: r^*\cong {\frac{r^{3}}{3\,Q^{2}},}  \label{r*-r-at-0}
1194: \end{equation}
1195: and hence \be V_{{\rm eff}}\cong -2{\frac{Q^{4}}{r^{6}}}\cong
1196: -{\frac{2}{9\,r^{*~2}}~.} \label{two-ninths}\ee
1197:  Thus even though the wave equation in the form (\ref{separation-of-var})
1198: is smooth, the effective potential in the normal form is unbounded
1199: from below; hence we will study later (in section
1200: \ref{cross-section-section}) its response to high-frequency waves,
1201: where all features of the potential other than its leading term
1202: $\propto r^{*~-2}$ are irrelevant.
1203:  Note that such potentials of the type $V=-c/r^{*~2}$ are special (see for instance \cite{LL35}).
1204: Considering a larger
1205:  class of potentials $V=-c/r^{*~\alpha}$ with $c,\alpha>0$, then $\alpha=2$ is a
1206: critical value describing a conformal quantum mechanics
1207:  \footnote{if we replace $\omega^2 \to \omega$ in (3.14).},
1208:  namely $c$ is dimensionless. For $0<\alpha<2$ the spectrum is bounded from
1209: below even though the  classical potential is not, while for
1210: $\alpha>2$ the spectrum is not bounded from below. For the
1211: critical value $\alpha=2$ there is a critical value for $c$,
1212: $c=1/4$. For $c<1/4$ the potential is still mild enough to have a
1213: spectrum bounded from below (actually at zero),
1214:  while for $c>1/4$ it is not. So the numerical coefficient $c=2/9$ in (\ref{two-ninths}) is
1215: important. We will see later (in section \ref{generalizations})
1216: that $c$ varies over our examples (it depends on the number of
1217: dimensions, but always $c<1/4$ in the charged case), and that it
1218: determines the high-energy transmission properties of the
1219: singularity.
1220: 
1221: At $r^{*}=0$ the radial equation (\ref{normal-radial}) has a
1222: regular singularity, and we may look for the characteristic
1223: exponents $\rho $ defined by $~\phi _{\omega }^{*}\cong r^{*~\rho
1224: }$. These satisfy the characteristic equation
1225: \[
1226: \rho \,(\rho -1)+{\frac{2}{9}=0,}
1227: \]
1228: leading to the two roots
1229: \begin{equation}
1230: \rho _{1,2}={\frac{1}{3}},\,{\frac{2}{3},}
1231: \end{equation}
1232: which are independent of $\omega $. Translating back from $r^{*}$
1233: to $r$ using the transformation (\ref{defn-g},\ref{r*-r-at-0}) we
1234: get the exponents $r^{0}\,$and $r^{1},$ exactly as expected for a
1235: smooth second-order equation.
1236: 
1237: For Schwarzschild one may repeat the calculations based on the
1238: leading behavior of $f_{{\rm Schw}}$ (\ref{fSchw-lead}) and find
1239: \be
1240:  r^{*}\cong -{\frac{r^{2}}{4M},} \ee leading to \be
1241:  V_{{\rm eff}}\cong -4{\frac{M^{2}}{r^{4}}}\cong -{\frac{1}{4\,r^{*~2}}.}
1242: \ee
1243:  Thus we have here the critical value $c=1/4$ (independent of the
1244: spacetime dimension, see section \ref{generalizations}).
1245: 
1246: Attempting $\phi _{\omega }^{*}\cong r^{*~\rho }$ one obtains the
1247: characteristic equation
1248: \[
1249: \rho \,(\rho -1)+{\frac{1}{4}=0,}
1250: \]
1251: which yields \be \rho _{1}=\rho _{2}={\frac{1}{2}.} \ee The
1252: equality of the two exponents implies the presence of a log term
1253: in one of these solutions. Again translating back the
1254: characteristic exponents to the $r$ coordinate we get the
1255: $r^{0},\,\log (r)$ behavior which was derived above from Eq. (\ref
1256: {Schw-var-sep}).
1257: 
1258: \subsection{Characteristic formulation}
1259: \label{char-subsection}
1260: 
1261: Having defined the ``tortoise'' coordinate $r^*$ (\ref{def-rstar})
1262: and the effective potential (\ref{Veff}) we can now return to
1263: spell out in detail the definition of time evolution in the
1264: characteristic formulation.
1265: 
1266: The coordinate $r^*$ approaches $-\infty $ as $r\to -\infty $ and
1267: $+\infty $ as $r\to r_{-}$. Notice that the effective potential
1268: decays as $r^{-2}\simeq $ $r^{*~-2}$ at large negative $r$, and as
1269: $f\propto r-r_{-}$ (which is exponentially small in $r^*$) at the
1270: inner horizon.
1271: Therefore, in each of these asymptotic boundaries the radial
1272: function is a superposition of the two asymptotic solutions
1273: \begin{equation}
1274: \phi^{*}_\omega \cong {\rm exp}(+ i\, \omega \,r^*) \, \mbox{  and  } \,
1275: \phi^{*}_\omega \cong {\rm exp}(- i\, \omega \,r^*)~.
1276: \label{gen-asymptotic}
1277: \end{equation}
1278: Defining the Eddington-like coordinates
1279: \[
1280: v:= t-r^*\,,\qquad u:= t+r^*,
1281: \]
1282: the above two asymptotic solutions at both edges $r^*\to \pm \infty $
1283: read, for a particular $\omega $ component of $\phi^*$,
1284: \[
1285: \phi^* \cong {\rm exp}(i\,\omega u)\,\, \mbox{  and  } \,\,\phi^*
1286: \cong {\rm exp}(i\,\omega v)~,
1287: \]
1288: respectively. The non-decomposed field at both asymptotic edges
1289: takes the form
1290: \[
1291: \phi^*(r^*\to \pm \infty )\cong f_{\pm }(v)+g_{\pm }(u),
1292: \]
1293: where $f_{\pm }(v)$ and $g_{\pm }(u)$ are arbitrary functions (the
1294: ``$\pm $'' stands for the two asymptotic regions $r^*\to \pm
1295: \infty $, i.e. to the inner horizon and to large negative $r$). In
1296: the characteristic initial-value formulation one needs to specify
1297: the ingoing component $f_{-}(v)$ at $\scri^-$ and the component
1298: $g_{+}(u)$ coming from the inner horizon. These two functions
1299: should uniquely determine the time evolution in the relevant piece
1300: of spacetime. This evolution is defined as follows. First,
1301: Fourier-decompose these two initial functions:
1302: \begin{eqnarray*}
1303: f_{-}(v) &=&\int f_{v}(\omega _{v})\,{\rm exp}(i\,\omega
1304: \,_{v}v)d\omega
1305: _{v}, \\
1306: g_{+}(u) &=&\int g_{u}(\omega _{u})\,{\rm exp}(i\,\omega
1307: \,_{u}u)d\omega _{u}.
1308: \end{eqnarray*}
1309: Next, we define a set of basis solutions $\phi _{(\omega
1310: _{v},0)}^{*}(r^{*},t)$ and $\phi _{(0,\omega _{u})}^{*}(r^{*},t)$
1311: (for each $\omega _{v}$ and $\omega_{u}$) as follows: $\phi
1312: _{(\omega _{v},0)}^{*}$ is the solution evolving from the
1313: characteristic initial conditions
1314: \[
1315: f_{-}(v)={\rm exp}(i\,\omega \,_{v}v)\,,\qquad g_{+}(u)=0.
1316: \]
1317: Similarly, $\phi _{(0,\omega _{u})}^{*}$ is the solution evolving
1318: from the characteristic initial conditions
1319: \[
1320: f_{-}(v)={\rm 0}\,,\qquad g_{+}(u)={\rm exp}(i\,\omega \,_{u}u).
1321: \]
1322: The two functions $\phi _{(\omega _{v},0)}^{*}$ and $\phi
1323: _{(0,\omega _{u})}^{*}$ are constructed from the radial functions
1324: $\phi _{\omega }^{*}(r^*)$ (with $\omega=\omega_v$ and
1325: $\omega=\omega_u$, respectively); see appendix \ref{char-appendix}
1326: for details.
1327: 
1328: 
1329: Once $\phi _{(\omega _{v},0)}^{*}$ and $\phi _{(0,\omega
1330: _{u})}^{*}$ are defined, the time evolution of the field $\phi^*$
1331: emerging from the characteristic initial data $f_{-}(v)$ and
1332: $g_{+}(u)$ is simply given by
1333: \begin{eqnarray*}
1334: \phi^*(r^*,t) &=&\int f_{v}(\omega _{v})\,\phi _{(\omega
1335: _{v},0)}^{*}(r^*,t)d\omega _{v} \\
1336: &&+\int g_{u}(\omega _{u})\,\phi _{(0,\omega
1337: _{u})}^{*}(r^*,t)d\omega _{u}.
1338: \end{eqnarray*}
1339: We conclude that the unique extension of the radial functions
1340: $\phi _{\omega }^{*}(r^*)$ across the $r=0$ singularity naturally
1341: leads to a unique time
1342: evolution for any set of characteristic initial functions $f_{-}(v)$ and
1343: $g_{+}(u)$.
1344: 
1345: 
1346: \section{Transmission cross section in \RN}
1347: 
1348: \label{cross-section-section}
1349: 
1350: 
1351: Having demonstrated a resolution of the $r=0$ singularity we turn
1352: in this section to examine its nature through scattering. Here we
1353: shall only consider the charged (Reissner-Nordstr\"{o}m) case.
1354: Full reflection from the negative-mass Schwarzschild singularity
1355: will be considered together with the $d>4$ cases at the end of
1356: subsection \ref{highd}. We focus here on the transmission of
1357: high-frequency incoming waves through the $r=0$ singularity (from
1358: now on we will use ``high-energy'' to mean ``high-frequency'' by a
1359: slight abuse of language originating from the quantum theory).
1360: 
1361:  We first study the effective potential $V_{{\rm
1362: eff}}$, and especially its asymptotic form at small $r$. Then we
1363: compute the transmission and reflection coefficients for fixed $l$
1364: in the high-energy limit. Finally we combine that result with the
1365: high-$l$ limit, where the partial cross section vanishes. We sum
1366: over $l$ and obtain the $\omega $ dependence of the (high-energy)
1367: total cross section for transmission.
1368: 
1369: \subsection{General features of the effective potential}
1370: 
1371: Consider first the qualitative behavior of the effective potential
1372: $V_{{\rm eff}}$. For concreteness we shall consider here a wave
1373: propagating from past null infinity of the negative-$r$ universe
1374: towards $r=0$. At large $|r|$, we have the usual centrifugal
1375: barrier
1376: \begin{equation}
1377: V_{{\rm eff}}\cong \frac{l(l+1)}{r^{2}}\qquad \qquad (|r|\gg M).
1378: \label{Vlarger}
1379: \end{equation}
1380: At small $|r|$, the potential takes the asymptotic form
1381: \begin{equation}
1382: V_{{\rm eff}}\cong
1383: l(l+1)\frac{Q^{2}}{r^{4}}-2\frac{Q^{4}}{r^{6}}\qquad \qquad
1384: (|r|\ll Q).  \label{Vsmallr}
1385: \end{equation}
1386: Here we only keep the leading order in $r/Q$. We do keep, however,
1387: the term $ \propto r^{-4}$ in the right-hand side because it is
1388: proportional to $l(l+1)$. This term will become important for
1389: partial waves of sufficiently large $l$, which are involved in the
1390: calculation of the total high-energy transmission cross-section
1391: (see below).
1392: 
1393: The contribution of any particular $l$ to the total high-energy
1394: cross-section is suppressed by the kinematic factor $k^{-2}$ in
1395: Eq. (\ref {total-cross-section}) below (while it will turn out
1396: that $\sigma \gg k^{-2}$). Therefore, at high energy the total
1397: cross section is dominated by partial waves with $l\gg 1$, which
1398: we shall assume hereafter. Such waves of relatively small or
1399: moderate energy will be reflected already at the large-$|r|$
1400: centrifugal barrier (\ref{Vlarger}). Only waves with sufficiently
1401: high energy, $\omega >l/Q,$ will penetrate into the central region
1402: $|r|\ll Q$. The scattering features of these high-energy waves
1403: will be dominated by the small-$|r|$ potential (\ref{Vsmallr}).
1404: This potential has a peak value (hereafter we often replace
1405: $l(l+1)$ by $l^{2}$, which is justified because $l\gg 1$)
1406: $$
1407: V_{\max }\cong \frac{l^{6}}{27Q^{2}},
1408: $$
1409: located at
1410: $$
1411: |r|\cong r_{peak}:= \sqrt{3}Q/l\,.
1412: $$
1413: It acts as a repulsive barrier ($\propto r^{-4}$) at $Q\gg
1414: |r|>r_{peak}$, and as a potential well ($\propto -r^{-6}$) at
1415: $|r|<r_{peak}$.
1416: 
1417: The above discussion immediately suggests that partial waves with
1418: $\omega \ll \omega _{peak}$, where
1419: \begin{equation}
1420: \omega _{peak}:= \sqrt{V_{\max }}\cong \frac{l^{3}}{\sqrt{27}Q}\,,
1421: \label{wpeak}
1422: \end{equation}
1423: will be fully reflected by the potential barrier at
1424: $|r|>r_{peak}$. However, partial waves with $\omega \gg \omega
1425: _{peak}$ will predominantly feel the potential well $\propto
1426: -r^{-6}$ at $|r|<r_{peak}$. These waves will be partially
1427: transmitted, as we analyze in the next subsection.
1428: 
1429: For later convenience we transform the small-$|r|$ potential from
1430: $r$ to $r^*$, using $r^{3}\cong 3\,Q^{2}\, r^*$:
1431: \begin{equation}
1432: V_{{\rm eff}}\cong
1433: \frac{l^{2}}{Q^{2/3}(3r^*)^{4/3}}-{\frac{2}{9\,r^{*~2}}} \qquad
1434: \qquad (r\ll Q).  \label{Vsmallr*}
1435: \end{equation}
1436: The peak is located at
1437: $$
1438: |r^*|\cong r^*_{peak}:= \sqrt{3}Q/l^{3}\,.
1439: $$
1440: Although we are primarily considering here a wave coming from the
1441: negative-$r$ past null infinity, the case of a wave propagating
1442: from $r_{-}$ towards $r=0$ may be treated in exactly the same
1443: manner. Note that the small-$|r|$ potential (\ref{Vsmallr}) or
1444: (\ref{Vsmallr*}) is even in $r$ or $r^*$, respectively.
1445: Consequently the high-energy transmission amplitude for partial
1446: waves is the same for both cases. (Obviously the transmission
1447: probability is exactly the same, even for finite $\omega $.)
1448: 
1449: \subsection{Amplitudes at high energy and fixed $l$}
1450: 
1451: \label{trasnmission-reflection}It is instructive to look at the
1452: scattering in the high-energy limit. Normally, the potential is
1453: bounded, hence the transmission amplitude approaches $1$ and the
1454: reflection vanishes at this limit. However, here we have an
1455: unbounded (attractive) potential near $r=r^{*}=0$, which leads to
1456: partial reflection even at the high-energy limit.
1457: 
1458: For fixed $l$ and high energy, $\omega \gg \omega _{peak}$, the
1459: potential is dominated by
1460: $$
1461: V_{{\rm eff}}\cong -{\frac{2}{9\,r^{*~2}}\equiv }V_{{\rm
1462: eff,sing}}\,.
1463: $$
1464: We therefore need to solve the scattering problem
1465: \begin{equation}
1466: 0=[-{\partial }_{r*r*}+V_{{\rm eff,sing}}(r^*)-\omega ^{2}]\phi
1467: ^{*}. \label{V-singular}
1468: \end{equation}
1469: Passing to the dimensionless variable
1470:  \footnote{The negative sign
1471: in the definition of $x$ was introduced in order for the incoming
1472: wave to propagate from large positive $x$ values (null infinity)
1473: towards smaller $x$ values.}
1474: \begin{equation}
1475: x=-r^{*}\,\omega ,  \label{xr*}
1476: \end{equation}
1477: the equation is transformed into
1478: \begin{equation}
1479: 0=[-{\partial }_{xx}-{\frac{2}{9\,x^{2}}}-1]\,\phi ^{*}.
1480: \end{equation}
1481: Its general solution is
1482: \begin{equation}
1483: \phi ^{*}=c_{1}\,\phi _{1}^{*}+c_{2}\,\phi _{2}^{*}\,,
1484: \label{high-energy-soln}
1485: \end{equation}
1486: with
1487: \begin{eqnarray}
1488: \phi _{1}^{*} &=&\sqrt{x}\,J_{1/6}(x),  \nonumber \\
1489: \phi _{2}^{*} &=&\sqrt{x}\,J_{-1/6}(x)  \label{positivex}
1490: \end{eqnarray}
1491: (for $x>0$), where $J_{n}$ are Bessel functions of order $n$ (see
1492: appendix \ref{BesselApp}). The coefficients $c_{1}\,,c_{2}$ will
1493: be determined below.
1494: 
1495: In order to extend the solution (\ref{high-energy-soln}) across
1496: $x=0$, we first express it in terms of $r$. Recall that at small
1497: $r$,
1498: $$
1499: r^*\cong {\frac{r^{3}}{3\,Q^{2}},}
1500: $$
1501: and, furthermore, $r^*(r)$ is analytic at $r=0$. From the local
1502: behavior of the Bessel functions $J_{n}(\lambda )$ near $\lambda
1503: =0$ (see appendix \ref{BesselApp}), and that of $r^*$ near $r=0$,
1504: it follows that $\phi _{1}^{*} $ and $\phi _{2}^{*}$ take the form
1505: \begin{eqnarray}
1506: \phi _{1}^{*} &=&r^{2}\,A_{1/6}^{J}(r^{6}),  \nonumber \\
1507: \phi _{2}^{*} &=&rA_{-1/6}^{J}(r^{6}),  \label{fir}
1508: \end{eqnarray}
1509: where $A_{\pm 1/6}^{J}(\lambda )$ are some real functions (for
1510: real $\lambda $), analytic at $\lambda =0$. Obviously both $\phi
1511: _{1}^{*}(r)$ and $\phi _{2}^{*}(r)$ have a unique real analytic
1512: extension through $r=0$.\footnote{Recall that $\phi _{1}^{*}(r)$
1513: and $\phi _{2}^{*}(r)$ must extend to $r>0$ as real and analytic
1514: functions, because the radial equation (\ref {separation-of-var})
1515: is real and analytic (and $\phi _{1}^{*},\phi _{2}^{*}$ are real
1516: at $r<0$).} The form of Eq. (\ref{fir}) ensures that $\phi
1517: _{1}^{*}$ extends as an even function, and $\phi _{2}^{*}$ as an
1518: odd one. Hence for negative $x$ we get
1519: \begin{eqnarray}
1520: \phi _{1}^{*}(x) &=&\sqrt{|x|}\,J_{1/6}(|x|),  \nonumber \\
1521: \phi _{2}^{*}(x) &=&-\sqrt{|x|}\,J_{-1/6}(|x|).  \label{negativex}
1522: \end{eqnarray}
1523: In order to get the coefficients $c_{1}\,,c_{2}$  --  and the
1524: transmission and reflection coefficients  --  we impose the
1525: large-$r^{*}$ boundary conditions. Since we are considering here a
1526: wave propagating from (say) past null infinity (large negative
1527: $r^*$), we assume that no waves are entering from the inner
1528: horizon $(r^{*}\to +\infty )$. This corresponds to pure asymptotic
1529: behavior $\propto {\rm exp}(-i\,\omega \,r^{*})={\rm exp}(i\,x)$
1530: at $x\to -\infty $. Thus, in terms of $x$ the asymptotic behavior
1531: at the boundaries is
1532: \begin{equation}
1533: \begin{array}{cclcc}
1534: \phi ^{*} & \cong  & {\rm exp}(i\,x)+R(\omega )\,{\rm exp}(-ix) &
1535: \mbox{  at
1536: } & x\to +\infty \,,\nonumber \\
1537: \phi ^{*} & \cong  & T(\omega )\,{\rm exp}(ix) & \mbox{  at  } &
1538: x\to - \infty \,.
1539: \end{array}
1540: \label{boundary}
1541: \end{equation}
1542: Note that since $\omega$ got scaled out of the equation
1543: $T(\omega),\, R(\omega)$ are actually independent of $\omega$.
1544: 
1545: The asymptotic form of the Bessel functions for large (positive)
1546: argument is given in Eq. (\ref{BesselAsymp}). For $n=1/6$ it reads
1547: $$
1548: {\rm exp}(\pm i|x|)\cong \sqrt{\pi \,|x|/2}\,{\rm exp}(\pm i\pi
1549: /3)\left[ (1 \pm i\sqrt{3})J_{1/6}(|x|)\mp 2iJ_{-1/6}(|x|)\right]
1550: .
1551: $$
1552: Applying it first to the inner-horizon boundary ($x\to -\infty $),
1553: along with Eq. (\ref{negativex}), one finds
1554: \begin{eqnarray*}
1555: c_{1} &=&\sqrt{\pi /2}\,{\rm exp}(-i\pi /3)(1-i\sqrt{3})T(\omega ), \\
1556: c_{2} &=&-i\sqrt{2\pi }\,{\rm exp}(-i\pi /3)T(\omega ).
1557: \end{eqnarray*}
1558: Similarly, when applied to the null-infinity boundary ($x\to
1559: +\infty $), along with Eq. (\ref{positivex}), it yields
1560: 
1561: \begin{eqnarray*}
1562: c_{1} &=&\sqrt{\pi \,/2}\left[ {\rm exp}(i\pi
1563: /3)(1+i\sqrt{3})+{\rm exp}(-i \pi /3)(1-i\sqrt{3})R(\omega )\right] , \\
1564: c_{2} &=&-i\sqrt{2\pi }\left[ {\rm exp}(i\pi /3)-{\rm exp}(-i\pi
1565: /3)R(\omega )\right] .
1566: \end{eqnarray*}
1567: Combining the two expressions for $c_{1}\,,c_{2}$ and solving for
1568: $R$ and $T$, one obtains \be
1569:  R=\frac{\sqrt{3}}{2}i\, , \qquad T=\frac{1}{2} \,.
1570:  \ee
1571: 
1572: We conclude that for any $l$, at the high-energy limit a fraction
1573: $|R|^{2}=3/4$ of the influx is reflected and a fraction
1574: $|T|^{2}=1/4$ is transmitted to the other side of the $r=0$
1575: singularity. As was mentioned above, this applies to both waves
1576: coming from past null infinity (of the negative-$r$ universe
1577: beyond the singularity) and waves coming from the inner horizon
1578: towards the $r=0$ singularity.
1579: 
1580: \subsection{Summing over $l$}
1581: 
1582: \label{summing-l}
1583: 
1584: 
1585: The total cross section for transmission is given by a sum over
1586: all partial waves $l$:
1587: \begin{equation}
1588: \sigma ={\frac{\pi }{k^{2}}}\sum_{l=0}^{\infty
1589: }(2\,l+1)\,|T_{l}|^{2}, \label{total-cross-section}
1590: \end{equation}
1591: where $k$ is the momentum of the incoming plane wave ($k=\omega $
1592: for a massless field, which we assume from now on in this
1593: subsection). We shall consider here the cross section at the
1594: high-energy limit. As mentioned above, at this limit the total
1595: cross section is dominated by the contribution from modes $l\gg
1596: 1$, because each particular $l$ is suppressed by the factor
1597: $\omega ^{-2}$, while it will turn out that $\sigma \gg \omega
1598: ^{-2}$.
1599:  The result of the previous subsection, $|T_{l}|^{2}\cong 1/4$,
1600: holds for $l$ values that are not too large. However, for any
1601: fixed and large $\omega $ there is some large enough $l$ where the
1602: transmission practically vanishes. For instance, for classical
1603: flat-space scattering off a rigid target of length scale $L$
1604: (``rigid'' means here $\omega $-independent $L$), if the impact
1605: parameter $b$ is much larger than $L $ then the absorption
1606: vanishes. Since $b=l/p$ (and $p=\omega $ in our massless case) we
1607: see that for $l>L\omega $ the absorption vanishes. We will now
1608: show that in our case the suppression of transmission already
1609: happens for $l>(Q\omega \,)^{1/3}$.
1610: 
1611: A rough estimate of the total large-$\omega$ cross section can be
1612: easily done based on the qualitative features of $V_{{\rm eff}}$
1613: at small $r$. For $\omega \gg \omega _{peak}$, where $\omega
1614: _{peak}$ is defined in (\ref{wpeak}), the above high-energy result
1615: $|T_{l}|^{2}\cong 1/4 $ holds. For $\omega $ significantly smaller
1616: than $\omega _{peak}$ the incoming wave will encounter the
1617: potential barrier \be
1618:  V_{{\rm eff}}\cong \frac{l^{2}}{Q^{2/3}(3r^*)^{4/3}}
1619:  \label{Vbarrier}
1620:   \ee and will be tunnelling suppressed there. [In fact, an
1621: effective tunnelling suppression requires that $\omega \Delta$ is
1622: greater than $1$, where $\Delta $ is the tunnelling length scale
1623: in the barrier in terms of $r^*$, i.e. from (\ref{Vbarrier})
1624: $\Delta \sim (\omega ^{3}Q/l^{3})^{-1/2}$. But this is
1625: automatically satisfied because $\omega \Delta \sim (\omega
1626: Q/l^{3})^{-1/2}$ is roughly $(\omega _{peak}/\omega )^{1/2}$,
1627: which is $\gg 1$ in the present discussion.] Thus, for a rough
1628: estimate of the large-$\omega $ cross section we may substitute
1629: $|T_{l}|^{2}\cong 1/4$ for $l\ll l_{c}$ and negligible
1630: $|T_{l}|^{2}$ for $l\gg l_{c}$, where \be
1631:  l_{c}:= \sqrt{3}(Q\omega )^{1/3} \ee
1632: can be read off the expression for $\omega_{peak}$ (\ref{wpeak}).
1633: 
1634: Equation (\ref{total-cross-section}) then yields the total cross
1635: section:
1636: \begin{equation}
1637: \sigma \sim \frac{l_{c}^{2}}{\omega ^{2}}\sim
1638: {\frac{Q^{2/3}}{\omega ^{4/3}}.}  \label{cross-section1}
1639: \end{equation}
1640: The total cross section may also be obtained from $l_c$ by a simple argument:
1641: the critical maximal impact parameter for
1642: transmission is $b_{\rm max} = l_c/ \omega$, and hence the total size
1643: (cross-section) of the target for transmission is
1644: \begin{equation}
1645: \sigma \sim b_{\rm max}^2 = \frac{l_{c}^{2}}{\omega ^{2}}\sim
1646: {\frac{Q^{2/3}}{\omega ^{4/3}}. }  \label{cross-section2}
1647: \end{equation}
1648: 
1649: In the above estimate we have not taken into account the
1650: contribution coming from the range $l\sim l_{c}$. A more rigorous
1651: derivation of the cross section
1652: (\ref{cross-section1},\ref{cross-section2}), which properly treats
1653: this range as well, may be done by analyzing the rescaling
1654: properties of the small-$r$ scattering problem, i.e.
1655: \begin{equation}
1656: \lbrack -{\partial }_{r^{*}\,r^{*}}+V_{{\rm {eff}}}-\omega
1657: ^{2}]\phi =0
1658: \end{equation}
1659: with the potential (\ref{Vsmallr*}). Transforming to a new
1660: dimensionless variable
1661: \begin{equation}
1662: x:=-{\frac{l^{3}}{\,Q}r^{*},}
1663: \end{equation}
1664: the equation becomes
1665: $$
1666: \lbrack -{\partial }_{x\,x}+\hat{V}(x)-\Omega ^{2}]\phi =0,
1667: $$
1668: where
1669: $$
1670: \hat{V}(x)=-{\frac{2}{9\,x^{2}}}+{\frac{1}{(3x)^{4/3}}}
1671: $$
1672: and
1673: $$
1674: \Omega :={\frac{Q}{l^{3}}\omega .}
1675: $$
1676: The boundary conditions at large $|x|$ are, in accordance with Eq.
1677: (\ref{boundary}),
1678: \begin{equation}
1679: \begin{array}{cclcc}
1680: \phi ^{*} & \cong  & {\rm exp}(i \Omega x)+R(\Omega )\,{\rm exp}(-i \Omega x)
1681: & \mbox{  at } & x\to +\infty \,,\nonumber \\
1682: \phi ^{*} & \cong  & T(\Omega )\,{\rm exp}(i \Omega x)
1683: & \mbox{  at  } & x\to -\infty \,.
1684: \end{array}
1685: \label{boundary1}
1686: \end{equation}
1687: Hence in this limit of large $\omega $ and large $l$ the
1688: transmission amplitude depends only on $\Omega$, i.e.
1689: \begin{equation}
1690: T=T({\frac{\omega \,Q}{l^{3}}}).
1691: \end{equation}
1692: The cross section is
1693: $$
1694: \sigma ={\frac{\pi }{\omega ^{2}}}\sum_{l=0}^{\infty
1695: }(2\,l+1)\,|T_{l}|^{2} \cong {\frac{2\pi }{\omega ^{2}}}\int
1696: \,l\,\,|T({\frac{\omega \,Q}{l^{3}}} )|^{2}dl,
1697: $$
1698: where we used $l\gg 1$ to replace $2\,l+1$ by $2\,l$ and the sum
1699: by an integral. Finally, transforming from $l$ to
1700: $\hat{l}:=l/(\omega \,Q)^{1/3}$ we find
1701: \begin{equation}
1702: \sigma \cong c{\frac{Q^{2/3}}{\omega ^{4/3}},}
1703: \end{equation}
1704: where
1705: $$
1706: c=2\pi \int \,\hat{l}\,\,|T(\hat{l}^{-3})|^{2}d\hat{l},
1707: $$
1708: in agreement with the above estimate (\ref{cross-section1}).
1709: 
1710: 
1711: Having found the total cross section for transmission, one may
1712: inquire about the total cross section for scattering. At large
1713: distances this is just like a Coulomb problem with a length scale
1714: $\,M$. Due to the infinite range of the gravitational force, we
1715: expect the cross section for scattering to be infinite
1716: for a massive field.
1717: However, for a massless field the scattering potential decays
1718: (after subtracting the standard centrifugal piece
1719: $l(l+1)/r^{*~2}$) as $\ln (r^*/M)/r^{*~3}$. This is a short-range
1720: potential, hence we expect a finite scattering cross-section in
1721: the massless case.
1722: 
1723: The peculiar $\omega$ dependence of the transmission cross section
1724: (\ref{cross-section1}) through the singularity can be compared
1725: with two other systems. For black holes with non-zero area the
1726: large $\omega$ absorption cross section through the event horizon
1727: tends to the geodesic cross section, and hence $\sigma \sim
1728: \omega^0$. For elementary particles, on the other hand, the
1729: high-energy scattering cross section $\sigma_{\rm scat} \sim
1730: \omega^{-2}$ which is the kinematical factor which multiplies the
1731: dimensionless scattering amplitude. So we see that this system
1732: lies between the two examples above.
1733: 
1734: Another interpretation issue is whether the partial
1735: transmission/partial scattering behavior at large $\omega$ should
1736: be interpreted as a beam splitter for single particles. The
1737: question is whether a wave packet can be made small enough so that
1738: it fits into a target of size $\sim \omega^{-2/3}$ -- otherwise it
1739: will not be fully split. Imagine that we set a ``lens'' (one which
1740: affects $\phi$ waves) at a distance $f \sim \omega^{0} \gg Q$ from
1741: the black hole. We know that the spot size at focus can be made as
1742: small as $\lambda\, f/D$, where $\lambda=2\, \pi/\omega$ is the
1743: wavelength and $D$ is the aperture size, and we here imagine that
1744: space is flat and we need to hit a target with the same cross
1745: section as the $r=0$ singularity. From $\lambda\, f/D \ll
1746: \omega^{-2/3}$ we find that we need $D \ge \omega^{-1/3}$. So such
1747: focusing can be readily achieved, even with a small size lens.
1748: This suggest that the singularity may indeed act as a beam
1749: splitter. Note, however, that in the above analysis we only
1750: considered incoming {\it plane waves} (or pure partial waves with
1751: well-defined $l$ values). In order to verify the beam-splitter
1752: behavior one needs to explicitly analyze the transmission
1753: amplitude in the geometrical configuration considered above (i.e.
1754: a narrow beem focused on a small region near $r=0$). This
1755: calculation is beyond the scope of the present paper.
1756: 
1757: \section{Generalizing the backgrounds}
1758: \label{generalizations}
1759: 
1760: \subsection{Higher dimensions}
1761: \label{highd}
1762: 
1763: 
1764: The background RN metric and gauge field in $d \ge 4$ dimensions
1765: are \cite{MyersPerry} \bea
1766:  ds^2 &=& -f\, dt^2 + f^{-1}\, dr^2 + r^2\, d\Omega_{d-2}^{~2} \non
1767:  A &=& -{Q \over r^{d-3}}\, dt \non
1768:  f &=& 1-{r_+^{d-3}+r_-^{d-3} \over r^{d-3}} + {(r_+\, r_-)^{d-3} \over
1769: r^{2\,(d-3)}}
1770:   \label{RNmetric2}
1771:  \eea
1772:  The mass and charge are \bea
1773:  M &=& {(d-2)\, \Omega_{d-2} \over 16\, \pi}( r_+^{d-3} + r_-^{d-3} ) \non
1774:  Q &=& {\rm const}\,\sqrt{r_+ \, r_-}^{~d-3} \eea
1775: where $\Omega_{d-2}$ is the area of the $d-2$ sphere, and we do
1776: not fix the constant in front of the charge which depends on the
1777: normalization chosen for the vector field. In order to get
1778: $d$-dimensional Schwarzschild one sets $r_-=0, ~ r_+=r_0$.
1779: 
1780: After separation of the time variable ($\del_{tt} \rightarrow
1781: -\omega^2$) the scalar wave equation becomes \be
1782:   \br  {1 \over f\, r^{d-2}}\, \del_r\, f\, r^{d-2}\, \del_r -
1783:  {l\, (l+d-3) \over f\, r^2} + f^{-2}\, \omega^2 \kt \,
1784:  \phiom =0 ~.
1785:  \label{SeparatedWaveEqd} \ee
1786: 
1787: Considering the charged case first, and comparing with the 4d case
1788: (\ref{rnreg}) one notes a change: the equation
1789: (\ref{SeparatedWaveEqd}) is not regular anymore at $r=0$ (since
1790: $f\, r^{d-2}$ is has a pole).
1791:  Actually it is regular-singular with leading behavior \be
1792:   \br -\del_{rr} + {d-4 \over r}\, \( 1+O \( r^{d-3} \) \)
1793:   \del_r + l\, (l+d-3)\, O \( r^{2(d-4)}\) + \omega^2\, O \( r^{4(d-3)} \) \kt\, \phiom =0~.
1794:   \label{reg-sing-dg4} \ee
1795: The two solutions of the characteristic equation are $r^0,\,
1796: r^{d-3}$. Since the difference of the exponents is integral it is
1797: possible {\it a priori} that the $r^0$ solution has a $\log$ piece
1798: of the following form $\phiom=r^0 + \dots+ \log(r)\, r^{d-3} +
1799: \dots $, however it is seen not to be the case by explicitly
1800: Taylor expanding $\phiom$ in the differential equation
1801: (\ref{reg-sing-dg4}).
1802: 
1803: For Schwarzschild, on the other hand, one finds the same leading
1804: behavior as in 4d \be
1805:  \br \del_{rr} + {1 \over r}\,  \del_r + \dots \kt \phiom=0 \ee
1806: and thus the leading behavior of the two independent solutions is
1807: still $r^0,\, \log(r)$.
1808: 
1809: As in subsection \ref{normal-form-subsection} we may find the
1810: normal form of the radial equation (\ref{SeparatedWaveEqd}). The
1811: $r^*$ coordinate is defined by \be
1812:  dr^* = {dr \over f} ~~,\ee
1813:  the field is redefined by \be
1814:  \phi^* := r^{(d-2)/2}\phi~, \ee
1815: and the resulting effective potential is
1816:  \bea V_{\rm eff} &=& \Delta V_{\rm eff} + f\, {l\, (l+d-3) \over r^2} \non
1817:    \Delta V_{\rm eff} &=& {d-2 \over 2}\, {f \over r^{(d-2)/2}} \del_r
1818: (r^{(d-4)/2}\, f)
1819:   \label{Veffd} \eea
1820: where we use the same notations as in subsection \ref{normal-form-subsection}.
1821: 
1822: We start by looking at the charged case. The leading behavior near
1823: the singularity (at $r=0$) is \bea
1824:  r^* &\cong& {r^{2\, d-5} \over (2 \, d -5)\, (r_+\, r_-)^{(d-3)}}  \non
1825:  V_{\rm eff} &\cong& -{(d-2)\, (3d-8) \over 4\, (2d-5)^2}\, {1
1826: \over r^{*~2} }   \label{Veff-sing-RN} \eea
1827:  Given an effective potential $V_{\rm eff} \simeq -c/r^{*~2}$ the order
1828: $n$ of the Bessel functions which appear in the high $\omega$
1829: solution is given by (see appendix \ref{BesselApp}) \be
1830:  n^2 = {1\ \over 4} - c ~~.\ee
1831: Hence, the second equation in (\ref{Veff-sing-RN}) determines \be
1832:  |n|_{\rm RN}={1 \over 4} \, {d-3 \over  d-5/2} \ee
1833:  generalizing the 4d result $|n|=1/6$ (\ref{positivex}).
1834: 
1835: We note in passing that for the 2d black hole we will show
1836: in the next subsection
1837: that $V_{\rm eff} \simeq -3/(16 r^{*~2})$ and hence one can define
1838: an ``effective'' or ``equivalent'' RN dimension to be $d_{\rm
1839: eff}= \infty$ or $11/4$, and $|n|_{2d}=1/4$.
1840: 
1841: The transmission and reflection coefficients for high $\omega$
1842:  and fixed $l$ follow from the asymptotics of the Bessel functions as in
1843: subsection \ref{trasnmission-reflection}:
1844:  \bea |R| &=&  |\cos (n\, \pi)| \non
1845:      |T| &=&    |\sin (n\, \pi)| \label{rrtt} \eea
1846: The $\omega$ dependence of the total cross section for
1847: transmission can be derived along the lines of subsection
1848: \ref{summing-l}. The leading terms in the potential (\ref{Veffd})
1849: at $r \sim 0$ are \be
1850:  V_{\rm eff} \simeq l^2\, {(r_+\, r_-)^{d-3} \over r^{2\, d-4}}-
1851:  {d-2 \over 2}\, ({3 \over 2}\, d -4)\, {(r_+\, r_-)^{2(d-3)} \over r^{4\,
1852:  d-10}} ~ .\ee
1853:  For fixed $l$ the potential has a maximum at \be
1854:  r_{peak} \simeq {1 \over \sqrt[d-3]{l}} ~.\ee
1855:  Equating $V(r_{peak})$ with $\omega^2$ we get the transition value
1856:  for $l$: \be
1857:  l_c \sim \omega^{d-3 \over 2\, d -5}~. \ee
1858: % r_{\mbox{max}} \sim {1 \over \omega^{2\, d-5}} \\
1859: Finally, we evaluate the total transmission cross section by
1860: employing qualitative classical-mechanics considerations.
1861: Classically the cross section is obtained by estimating the
1862: maximal value of impact parameter $b$ for which a significant
1863: transmission (or scattering, in the more general context) still
1864: occurs. Using $b=l/\omega$ we get $b_{\rm max} \sim l_c/\omega$.
1865: Since in $d$ dimensions the cross section scales as $b_{\rm
1866: max}^{d-2}$, we find \be
1867:  \sigma \sim b_{\rm max}^{d-2} \sim \left( {l_c \over \omega}
1868:  \right)^{d-2} \sim \omega^{-{(d-2)^2 \over 2\, d-5}}~. \ee
1869: 
1870: 
1871: For \Schw black holes the leading behavior near the singularity
1872: (at $r=0$) is \bea
1873:  r^* &\cong& -{r^{d-2} \over (d -2)\, r_0^{d-3}}  \non
1874: V_{\rm eff} &\cong& -{1 \over 4\, r^{*~2} }
1875:     \label{Veff-sing-Schw} ~.\eea
1876: Thus independently of dimension one has $c=1/4$ and hence total
1877: reflection: \bea
1878:  n &=& 0 \non
1879:  T &=& 0 ~~.\eea
1880: 
1881: \subsection{The 2d black hole}
1882: \label{tdss}
1883: 
1884: Black holes in two dimensions are apparently outside the pattern
1885: of the $d\geq 4$ black holes discussed in the previous sections
1886: (although as we shall recall \cite{GiveonRabinoviciSever} they
1887: turn out to share similar features). Such solutions require, for
1888: instance, the presence of a dilaton. In this subsection we shall
1889: consider two dimensional dilaton gravity with an Abelian gauge
1890: field which is inspired by string theory. The action
1891: is~\footnote{This can be obtained, for instance, as (part of) a
1892: low-energy effective action of an heterotic string in two
1893: dimensions (for a review, see \cite{Polchinski}).}
1894:  \be
1895:    S=\int d^2 x\sqrt{-g}e^{-2\Phi}\left(
1896:    R+4g^{\mu\nu}\del_\mu\Phi\del_\nu\Phi-{1\over 4} F^2-\lambda\right)~,
1897:  \ee
1898: where $g_{\mu\nu}$ is the (string frame) metric, $\mu,\nu=0,1$,
1899: $\Phi$ is the dilaton, $F_{\mu\nu}$ is the field strength of an
1900: Abelian gauge field $A_{\mu}$, and $\lambda$ is the cosmological
1901: constant. In this two dimensional theory, the charged black hole
1902: solution is given by \cite{nappiyost}
1903:  \bea
1904:     ds^2 &\propto & f^{-1}\, d\rho^2 -f\, dt^2 \non
1905:     A &\propto& -{Q\over r}\, dt \non
1906:     \Phi(\rho) &=& \Phi_0- \half \rho~, \label{twodsol}
1907:  \eea
1908: where $\Phi_0$ is a constant and
1909:  \be
1910:     f(r)=1-{2M\over r}+{Q^2\over r^2}\ ,\qquad r=e^\rho~.
1911:     \label{twodf}
1912:  \ee
1913: Note that here $ds^2\propto f^{-1}\, {dr^2\over r^2} -f\, dt^2$,
1914: which is different from the generic $d\geq 4$ cases in eq.
1915: (\ref{RNmetric2}). Nevertheless, the geometry of this 2d
1916: black-hole is similar to a two dimensional slice of the
1917: Reissner-Nordstr\"{o}m solution, whose complete Penrose diagram is
1918: given in figure \ref{PenroseRN}. In the 4d case every point in
1919: figure \ref{PenroseRN} is actually a two sphere, while in the 2d
1920: case there is a non-trivial dilaton instead.
1921: 
1922: 
1923: An uncharged scalar field is minimally coupled to the background
1924: above as follows: \be
1925:    S=\int d^2 x\sqrt{-g}e^{-2\Phi}\left(
1926:    R+4g^{\mu\nu}\del_\mu\Phi\del_\nu\Phi-{1\over 4} F^2-\lambda -
1927:    g^{\mu\nu}\del_\mu \psi\del_\nu \psi-m^2\psi^2\right)~,
1928:  \ee
1929: where in this section
1930: $\psi$ denotes the scalar field to distinguish it
1931: from the dilaton $\Phi$. The wave equation of $\psi$ is:
1932:  \be
1933:     \Box_\Phi\psi={e^{2\Phi}\over\sqrt{-g}}\, \del_\mu\,
1934:     \sqrt{-g}\, e^{-2\Phi}\, g^{\mu\nu}\,
1935:     \del_\nu\psi=\(\del_rr^2f\del_r-f^{-1}\del_{tt}\)\psi=
1936:     m^2\psi~, \label{laplace}
1937:  \ee
1938: where we used $g_{rr}=1/(f\, r^2)$ and $\sqrt{-g}=e^{2(\Phi -\Phi_0)}=1/r$.
1939: Hence, in the two dimensional case the dilaton $\Phi$ plays again
1940: the role of the 4d spherical coordinates, now in the wave
1941: equation.
1942: 
1943: As in the $d\geq 4$ cases, we separate variables
1944:  \be
1945:     \psi(r,t)=\psi_\omega(r)e^{i\omega t}~.
1946:  \ee
1947: The equation for $\psi_\omega$ is:
1948:  \be
1949:     \(     \del_rr^2f\del_r+\omega^2f^{-1}-m^2\)\psi_\omega = 0~.
1950:     \label{eqpsi}
1951:  \ee
1952: Again, we define a coordinate $r^*$,
1953:  \be
1954:     dr^*={dr\over rf(r)}~,
1955:  \ee
1956: and a new field variable
1957:  \be
1958:     \psi^{*}:= \sqrt{r}\psi~,
1959:  \ee
1960: such that eq. (\ref{eqpsi}) turns into:
1961:  \be
1962:     \left[-\del_{r^*r^*}+V_{{\rm eff}}(r^*)-\omega^2\right]\psi^{*}_\omega=0~.
1963:  \label{turnsin}
1964:  \ee
1965: The effective potential is now
1966:  \be
1967:     V_{{\rm eff}}(r^*)=\Delta V_{{\rm eff}}+fm^2~,
1968:  \ee
1969: where
1970:  \be
1971:     \Delta V_{{\rm eff}}=\half\sqrt{r}f\del_r(\sqrt{r}f)~.
1972:  \ee
1973: Next we inspect the behavior of the wave equation (\ref{eqpsi})
1974: near the singularity of the 2d RN-like black hole. When $Q\neq 0$:
1975:  \bea
1976:     &{1\over Q^2}\(\del_r r^2f\del_r+\omega^2f^{-1}-m^2\)\psi_\omega=
1977:     \non &\left[(1+...)\del_{rr} +O(r^0)\del_r-{m^2\over Q^2}+
1978:     (1+...){r^2\over Q^4}w^2\right]\psi_\omega = 0~,
1979:  \eea
1980: where ``$...$'' stand for higher orders in $r$. As in the 4d case
1981: (\ref{rnreg}), this equation is regular at $r=0$ for every
1982: $\omega$. Moreover, near $r\sim 0$ we have
1983:  \be
1984:     r^*\cong{r^2\over 2Q^2}
1985:  \ee
1986: and hence
1987:  \be
1988:     V_{{\rm eff}}(r^*)\cong -{3\over 16\, r^{*~2}}\label{veff}~.
1989:  \ee
1990: On the other hand, for $Q=0$ -- the Schwarzschild-like 2d black
1991: hole -- the equation for $\psi_\omega$ is:
1992:  \bea
1993:     &-{1\over 2Mr}\(\del_rr^2f\del_r+\omega^2f^{-1}-m^2\)\psi_\omega=
1994:     \non &\left[(1+...)\del_{rr} +{(1+...)\over r}\del_r+{m^2\over 2Mr}+
1995:     (1+...){\omega^2\over 4M^2}\right]\psi_\omega = 0~.
1996:  \eea
1997: Again, as in the 4d case (\ref{Schw-var-sep}), it is not regular
1998: at the singularity, but instead it is regular-singular. In the
1999: uncharged case, to leading order near $r\sim 0$ we have
2000:  \be
2001:     r^*\cong -{r\over 2M}
2002:  \ee
2003: and hence
2004:  \be
2005:     V_{\rm eff}(r^*) \cong -{1 \over 4\, r^{*~2}}~.\label{veffunch}
2006:  \ee
2007: As discussed in previous sections, the solutions to high frequency
2008: scattering wave equations (\ref{turnsin}) are given in terms of
2009: Bessel functions, which depend on the effective potential. In the
2010: charged case we read the order of the relevant Bessel function
2011: from (\ref{veff}) and appendix \ref{BesselApp}: \be |n|={1 \over
2012: 4}~, \ee and hence at high energies we find (\ref{rrtt}):
2013:  \be
2014:     |T| = |R| = {1 \over \sqrt{2}}~. \label{rt}
2015:  \ee
2016: On the other hand, for the uncharged case, from (\ref{veffunch})
2017: and (\ref{rrtt}) we read:
2018:  \bea
2019:     n &=& 0 \non
2020:     |R| = 1&,& \quad T = 0~. \label{rtunch}
2021:  \eea
2022: To summarize, we see that the features of the 2d black holes
2023: discussed in this subsection are very similar to the 4d case.
2024: 
2025: 
2026: The 2d background (\ref{twodsol}) is also an exact Conformal Field
2027: Theory (CFT) background in string theory; in
2028: \cite{GiveonRabinoviciSever} it was obtained from a family of
2029: ${SL(2,\IR)\times U(1)\over U(1)}$ quotient CFT sigma models by a
2030: Kaluza-Klein reduction\footnote{In the bosonic case it is a
2031: semi-classical approximation, while in the superconformal
2032: extension this background is claimed to be exact \cite{bars}.}. In
2033: string theory one is forced to include the regions beyond the
2034: singularity \cite{giveon} (this is argued, for instance, by using
2035: T-duality; for a review, see \cite{gpr}). The structure of the
2036: parent $SL(2,\IR)$ group allows one \cite{GiveonRabinoviciSever}
2037: to find the exact solutions to the wave
2038: equation\footnote{Scattering waves are determined from vertex
2039: operators in the $SL(2,\IR)$ CFT. Those are given by matrix
2040: elements in a unitary representation of the group. Equivalently,
2041: these are solutions to the Laplace equation (\ref{laplace}), which
2042: in this case is a solvable hypergeometric equation.}. In
2043: particular, it gives the exact reflection coefficient
2044:  \be
2045:     |R(\omega)|^2={\ch\({r_-\omega \over r_+-r_-}\)\ch(\omega)
2046: \over\ch\({r_+\omega \over r_+-r_-}\)}~,
2047:  \ee
2048: where $\omega$ is proportional to the energy of a massless
2049: particle scattered from the region beyond the black hole
2050: singularity, and $r_{\pm}$ are the locations of the event and
2051: Cauchy horizons:
2052:  \be
2053:    r_\pm=M\pm\sqrt{M^2-Q^2}~.
2054:  \ee
2055: Indeed, for the charged case $|R|^2 \longrightarrow \half$ in the
2056: limit $\omega\rightarrow\infty$, and for the uncharged case
2057: $|R|^2=1$ for every energy. These results are in agreement with
2058: (\ref{rt}) and (\ref{rtunch}), respectively.
2059: 
2060: \section{Adding perturbations}
2061:  \label{perturbations-section}
2062: 
2063: 
2064: Until now we assumed that $\phi$ was a scalar field with no mass,
2065: no charge, no back-reaction and no other interactions beyond
2066: minimal coupling to gravity. Now we will check whether our result
2067: that the time evolution is well-defined is disrupted  by any of
2068: these perturbations. To our surprise we do not encounter any
2069: problem.
2070: 
2071: 
2072: \subsection{Mass and charge}
2073: 
2074: 
2075: Adding a mass $m$ and charge $q$ for the field $\phi$ is
2076: completely captured by the form of the geodesic potential
2077: (\ref{Veff-geod}), namely one has \bea
2078:  V_{\rm eff}^{\rm wave} &=& \tilde{V}_{\rm eff}^{\rm geodesic} + \Delta
2079: V_{\rm eff} \non
2080:  \tilde{V}_{\rm eff}^{\rm geodesic} &=&
2081: \left({\frac{l\,(l+d-3)}{r^{2}}}\,+m^{2}\right)\,f(r)
2082:   -\left(E-{\frac{q\,Q}{r^{d-3}}}\right)^{2} \non
2083:  \Delta V_{\rm eff} &=& {d-2 \over 2}\, {f \over r^{(d-2)/2}} \del_r
2084: (r^{(d-4)/2}\, f)
2085:  \eea
2086:  where $\tilde{V}_{\rm eff}^{\rm geodesic}$ is simply
2087:  $V_{\rm eff}^{\rm geodesic}$ from (\ref{Veff-geod}) with the
2088:  substitutions
2089:  $l^2 \to l(l+d-3)$
2090: and $E \to \omega$,
2091: and $\Delta V_{\rm eff}$ is the same as in the zero-mass
2092: zero-charge case (\ref{DeltaVeff},\ref{Veffd}).
2093: 
2094: Our analysis relied on the leading terms near $r \sim 0$, which we
2095: will now find to dominate over the added perturbations. For $Q
2096: \neq 0$ these are $1/r^{4 d -10}$ from $\Delta V_{\rm eff}$ and
2097: $l^2/r^{2d-4}$ from $\tilde{V}_{\rm eff}^{\rm geodesic}$. Since
2098: both $m$ and $q$ contribute at order $1/r^{2d-6}$, they are
2099: irrelevant near the singularity (for $d \ge 4$). Moreover, one can
2100: confirm that the form of the subleading terms still conform with
2101: (\ref{reg-sing-dg4}) and hence there are no log pieces in the
2102: solutions.
2103: 
2104: Note however that since for $q \neq 0$ one cannot solve for
2105: $\omega^2$ in the wave equation and view it as eigen-values of
2106: some operator, any operator approach and considerations of
2107: self-adjointness (see section \ref{conditions-section}) would need
2108: to be changed.
2109: 
2110: Finally, for $Q=0$ we might as well take $q=0$ and the leading
2111: terms are $1/r^{2 d -4}$ from $\Delta V_{\rm eff}$ and
2112: $l^2/r^{d-1}$ from $\tilde{V}_{\rm eff}^{\rm geodesic}$, while $m$
2113: contributes at $1/r^{d-3}$ and is again irrelevant.
2114: 
2115: \subsection{Interactions}
2116: 
2117: We may add non-quadratic terms to the action, namely adding terms
2118: non-linear in $\phi$ to the wave equation. Much of our previous
2119: analysis, especially the separation of variables relied on
2120: linearity, hence we should reconsider it, starting from the
2121: non-linear wave equation \be
2122:  0=\Box \phi - g\, U(\phi) ~, \ee
2123:  where $g$ is a coupling constant and $U(\phi)$ is a non-linear
2124: potential term. One can attempt to solve this equation by
2125: perturbation theory with $g$ being the small parameter, namely one
2126: expands \be
2127:  \phi = \sum_j g^j\, \phi^{(j)} ~.\ee
2128: We start at zeroth order with a solution $\phi^{(0)}$ to the
2129: linearized equation \be
2130:  \Box \phi^{(0)}=0 \ee
2131: and consider whether the first correction is regular at $r=0$. The
2132: first correction satisfies \be
2133:  \Box \phi^{(1)} = g\, U(\phi^{(0)}) ~. \label{interactions-first-order} \ee
2134:  In order to solve this equation we separate the angular and time
2135:  variables \be
2136:  f^{-1}\, \Box_{\omega l}\phi^{(1)}_{\omega l} = g\, f^{-1}\, \left[
2137: U(\phi^{(0)})
2138:  \right]_{\omega l} ~. \ee
2139:  Since $\phi^{(0)}$ is regular by assumption, so are the
2140: components  $\left[ U(\phi^{(0)})
2141:  \right]_{\omega l}$. This is a non-homogeneous ordinary (second order)
2142: differential equation. A solution in the vicinity of the
2143: singularity may be obtained by Taylor expanding the equation (see
2144: also section \ref{conditions-section}). The homogeneous solutions
2145: are the smooth $\phi^{(0)}_{\omega l}$,
2146:  while the non-homogenous term added to (\ref{SeparatedWaveEqd}) is
2147: $O\(f^{-1}(r)\)$ and hence it is subleading and does not change
2148: our ability to obtain a (univalued) solution. This will happen at
2149: higher orders in $g$ as well.
2150: 
2151: 
2152: \subsection{Back-reaction for RN in $d \ge 4$}
2153: 
2154: So far we worked in the linear approximation where one neglects
2155: the back-reaction of the scalar field on the background geometry.
2156: Let us check whether including the back-reaction disrupts our
2157: result, namely whether we can find any obstruction for the
2158: solutions to the linear equation from being extended to solutions
2159: for the full non-linear system of background plus scalar field.
2160: Our analysis will be limited to lowest order back-reaction, and we
2161: shall not cover all cases, but the indications are that these
2162: solutions do survive back-reaction.
2163: 
2164: Let us define a small parameter $\eps$ such that $\phi$ is first
2165: order in $\eps$. Then at second order the background may change --
2166: the metric due to a $T_{\mu\nu}$ source, and the electromagnetic
2167: (EM) field due to the charge associated with $\phi$. For
2168: simplicity we may consider $\phi$ to be neutral. Since the metric
2169: is singular to start with, we cannot use the straight-forward
2170: criterion that the background remains regular after back-reaction.
2171: A conservative view would be to continue and look at third order
2172: whether the changes in $\phi$ as a result of the changes in the
2173: background make it singular. Another approach is to require that
2174: the singularity in the metric does not become worse after
2175: back-reaction. The latter has the advantage that one can stop at
2176: second order.
2177: 
2178: Moreover, for a charged black hole we can even save us the work of
2179: computing the second order by the following observation. Let us
2180: compare the stress-energy tensor of the scalar field with that of
2181: the EM field, namely the background source.
2182: The action for gravity + EM + an uncharged scalar field is
2183:  \be
2184:    S=\int d^d x\sqrt{-g} \left(
2185:    R-{1\over 4}F^2- g^{\mu\nu}\del_\mu \phi\del_\nu
2186:    \phi-m^2\phi^2 \right)~,
2187:  \ee
2188: where $g_{\mu\nu}$ is the metric, $F$ the field strength and
2189: $\phi$ an uncharged scalar field of mass $m$. The stress tensors
2190: of the gauge and scalar fields are
2191:  \bea
2192:    T_{\mu\nu}^{(F)}&=&\half F_{\mu\rho}F_{\nu\sigma}g^{\rho\sigma}-{1\over
2193:    8}g_{\mu\nu}F^2 \non
2194:    T_{\mu\nu}^{(\phi)}&=&\del_\mu\phi\del_\nu\phi- \half
2195:    g_{\mu\nu}\left[(\del\phi)^2+m^2\phi^2\right] \label{stresst}
2196:  \eea
2197: We start by comparing the scalar quantities \bea
2198:  (\del \phi)^2 &=& \del_r \phi\, \del_r \phi\ g^{rr} \sim  {1 \over
2199: r^{2(d-3)}} \non
2200:  (F^2)  &=& F_{rt}\, F_{rt}\, g^{rr}\, g^{tt} \sim {1 \over
2201:  r^{2(d-2)}} \eea
2202:  where we use the background (\ref{RNmetric2}) and the property that
2203: $\phi,\, \del_r\phi$ are regular at the singularity. We see that
2204: close to the singularity the scalar field seems to add a
2205: negligible source on top of the one from the EM field.
2206: 
2207: We now proceed to compare all components
2208:  \be \begin{array}{ll}
2209:  T_{rr}^{(F)} \sim  {1\over r^2}    \qquad     &    T_{rr}^{(\phi)} \sim 1 \\
2210:  T_{tt}^{(F)} \sim  {1\over r^{2(2d-5)}} \qquad &    T_{tt}^{(\phi)} \sim
2211: {1\over r^{4(d-3)}} \\
2212:  T_{rt}^{(F)}=0                          \qquad  &    T_{rt}^{(\phi)} \sim 1 \\
2213:  T_{\theta \theta}^{(F)} \sim  {1\over r^{2(d-3)}} \qquad &    T_{\theta
2214: \theta}^{(\phi)} \sim {1\over r^{2(d-4)}} \\
2215: \end{array}
2216:  \ee
2217: We see that indeed for the $rr,~tt$ and $\theta \theta$ components
2218: $T_{\mu \nu}^{(\phi)} \sim r^2 \, T_{\mu \nu}^{(F)}$ is
2219: negligible, while $T_{rt}^{(\phi)}$ is regular. Altogether we
2220: consider this to be strong evidence that back-reaction is weak and
2221: does not change the smoothness of the solutions.
2222: 
2223: For a charged field $\phi$ the simple argument above does not seem
2224: to work. In principle one should determine the back-reaction to
2225: the metric and gauge field and whether they are more or less
2226: singular than the original background. A preliminary analysis
2227: turned out to be involved, and so we did not reach any conclusions
2228: for this case. For similar reasons we do not discuss the
2229: back-reaction to the negative-mass \Schw either.
2230: 
2231: 
2232: \subsection{Back-reaction for the 2d black-hole}
2233: In two dimensions, the action for an uncharged scalar field
2234: coupled to the metric and dilaton is (see subsection \ref{tdss}):
2235:  \be
2236:    S=\int d^2 x\sqrt{-g}e^{-2\Phi}\(
2237:    R+4g^{\mu\nu}\del_\mu\Phi\del_\nu\Phi-{1\over 4} F^2-\lambda -
2238:    g^{\mu\nu}\del_\mu \psi\del_\nu \psi-m^2\psi^2\right)~,
2239:  \ee
2240: where $g_{\mu\nu}$ is the string frame metric, $\Phi$ the dilaton,
2241: $F$ is the field strength of an Abelian gauge field $A$, $\psi$ is
2242: the scalar field (to distinguish it from the dilaton) and
2243: $\lambda$ is the cosmological constant.
2244: 
2245: 
2246: Using the background (\ref{twodsol}) and the property that
2247: $\psi,\, \del_r\psi$ are regular at the singularity, we find the
2248: following behavior of the scalar stress tensor (\ref{stresst})
2249: near the singularity
2250:  \bea
2251:     e^{2\Phi}T_{rr}^{(\psi)}&\sim&1 \non
2252:     e^{2\Phi}T_{tt}^{(\psi)}&\sim&{1\over r^2} \non
2253:     e^{2\Phi}T_{rt}^{(\psi)}&\sim&1
2254:  \eea
2255: The dilaton stress tensor is
2256:  \be
2257:    e^{2\Phi}T_{\mu\nu}^{(\Phi)}=4\nabla_\mu\nabla_\nu\Phi
2258:    -4g_{\mu\nu}\left(\nabla^2\Phi-(\partial\Phi)^2-{1\over 4}\lambda\right)
2259:  \ee
2260: and has the following leading behavior near the singularity
2261:  \bea
2262:     e^{2\Phi}T_{rr}^{(\Phi)}&\sim&{1\over r^2} \non
2263:     e^{2\Phi}T_{tt}^{(\Phi)}&\sim&{1\over r^4} \non
2264:     e^{2\Phi}T_{rt}^{(\Phi)}&=0
2265:  \eea
2266: Again, there is no strong back-reaction to the metric, since
2267: component-wise each component of $T_{\mu\nu}^{(\psi)}$ is either
2268: subleading to $T_{\mu\nu}^{(\Phi)}$ or regular.
2269: 
2270: To check the back-reaction to the dilaton, we define
2271:  \be
2272:     \chi=e^{-\Phi} ~~,
2273:  \ee
2274: in terms of which the action becomes
2275:  \be
2276:    S=\int d^2 x\sqrt{-g}\br4g^{\mu\nu}\del_\mu\chi\del_\nu\chi+\chi^2
2277: \(R-{1\over 4} F^2-\lambda -
2278:    g^{\mu\nu}\del_\mu \psi\del_\nu \psi-m^2\psi^2\)\kt~.
2279:  \ee
2280: The equation of motion for $\chi$ is
2281:  \be
2282:     4\Box\chi-\(R-{1\over 4} F^2-\lambda -
2283:     g^{\mu\nu}\del_\mu \psi\del_\nu \psi-m^2\psi^2\)\chi=0~,
2284:     \label{eom}
2285:  \ee
2286: where $\Box$ is the Laplacian without the dilaton. In principle,
2287: we should expand the solution to (\ref{eom}) to second order in
2288: the perturbation and compare the correction to its initial
2289: value~\footnote{At zero order we have $\chi=\chi_0\sqrt{r}$
2290: (\ref{twodsol}, \ref{twodf}). The second solution to (\ref{eom})
2291: is subleading near $r=0$, so turning it on at higher orders would
2292: not change the result.}. However, just like for the metric, we
2293: recognize that the new source for $\chi$, ~($g^{\mu\nu}\del_\mu
2294: \psi\del_\nu \psi-m^2\psi^2$), is negligible at the singularity
2295: relative to the background sources ($R-{1\over 4} F^2$).
2296: Therefore, we do not expect a strong back-reaction for the dilaton
2297: either. Summarizing, we find indications for a weak back-reaction
2298: near the singularity of the 2d charged black-hole.
2299: 
2300: 
2301: 
2302: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2303: 
2304: \section{General conditions for resolution of a singularity}
2305: \label{conditions-section}
2306: 
2307: \subsection{Uniqueness of decomposition and natural boundary conditions}
2308: 
2309: In order for the ``singularity smoothing'' procedure which we
2310: defined in section \ref{smooth-section}  to make sense we need the
2311: existence and uniqueness of decomposition of any initial condition
2312: into a linear combination of eigen-functions $\phiom$. We know
2313: that Hermitian operators have this property, namely their
2314: eigen-functions form a basis of function space. Thus we are
2315: interested in representing the radial equation $\Box_{\omega l}\,
2316: \phiom=0$ through an operator and studying its Hermiticity, and so
2317: we define the operator $L$ by
2318:  \bea
2319:  f\, \Box_{\omega l} &=& \omega^2 - L[\del_r,r] \non
2320:  L &:=& -{f \over r^2}\, \del_r\, f\, r^2\, \del_r +
2321:  {f\, l\, (l+1) \over r^2}  \label{defL} \eea
2322: and we specialize back to 4d for definiteness. We note that $L$ is
2323: not regular at $r=0$ even for RN but only due to an overall factor
2324: of $f^2$ relative to the regular radial equation $f^{-1}\,
2325: \Box_{\omega l}$ (\ref{separation-of-var}). One may wonder whether
2326: the decomposition property would nevertheless hold for $L$. This
2327: will be answered in the negative as we shall show that the
2328: functions $\phiom$ do not span all possible functions.
2329: 
2330: Since the differential equation (\ref{separation-of-var}) is
2331: smooth at $r=0$ one can Taylor expand the equation and solutions
2332: \be \phi_\omega(r) = \sum_{j=0}^{\infty}\, \phi_\omega^{(j)}\, r^j
2333: \label{Taylor-phi} ~.\ee
2334:  Substituting this in  the differential equation transforms it into a
2335: recurrence equation for the Taylor coefficients
2336: $\phi_\omega^{(j)}$.
2337: 
2338: For RN one notes that the coefficient of $\omega^2$ is $f^{-2}$
2339: which is $O(r^4)$ and hence the recurrence equation is
2340: $\omega$-independent up to order $O(r^3)$ (inclusive) where
2341: $\phi_\omega^{(5)}$ is determined. Among the first 6 Taylor
2342: coefficients $\phi_\omega^{(0)}, \dots, \phi_\omega^{(5)}$ the
2343: first two may be considered to be initial conditions for the
2344: recurrence equation and the other four may be considered to be a
2345: function of them. In other words we find 4 $\omega$-independent
2346: linear relations among $\phi_\omega^{(0)}, \dots,
2347: \phi_\omega^{(5)}$. Owing to the $\omega$-independence of these
2348: relations they continue to hold for any function which can be
2349: expressed as a linear combination \be
2350:  \phi_i= \int\, d\omega\, \, \( \phi_+(\omega) + \phi_-(\omega) \)\,
2351:  \phi_{\omega}(r)~,
2352: \label{linear-combination} \ee
2353:  where $\phi_\pm(\omega)$ are arbitrary coefficients.
2354: This means that there are 4 conditions on the first 6 coefficients
2355: of any function in span($\phiom$).
2356: 
2357: For \Schw the phenomenon is similar but the details are different.
2358: We have $f^{-2}= O(r^2)$ and hence the recurrence relation is free
2359: up to (and including) the 4th coefficient $\phi_\omega^{(3)}$.
2360: Since we consider only the regular solutions then the recurrence
2361: relation requires a single initial condition $\phiom^{(0)}$  while
2362: $\phi_\omega^{(1)}, \dots, \phi_\omega^{(3)}$ may be considered to
2363: be an $\omega$-independent function of it, and hence there are
2364: necessarily 3 conditions on the first 4 Taylor coefficients of any
2365: function in span($\phiom$).
2366: 
2367: We see that ${\rm span}(\phiom)$ does not contain all functions,
2368: but rather there are constraints on the initial conditions at
2369: $r=0$. This is why we required the time-evolution to be
2370: well-defined only for wave packets prepared far away, but
2371: otherwise arbitrary. We expect that once the domain of $L$ is
2372: correctly defined it will be Hermitian and uniqueness of
2373: decomposition will hold and consequently the uniqueness of time
2374: evolution. The mathematical term that we expect to hold is that
2375: $L$ is ``essentially self-adjoint'' namely that it has a unique
2376: Hermitian (i.e. self-adjoint) extension (see \cite{ReedSimon} for
2377: a text book and \cite{HorowitzMarolf,IshibashiHosoya}).
2378: 
2379: 
2380: \sbsection{Initial conditions at r=0}
2381: 
2382: Let us find some necessary boundary conditions at $r=0$ in
2383: response to the ``non-span'' property just discussed.
2384: Since $L$ defined in (\ref{defL}) is singular  at $r=0$
2385: $\del_{tt}$ may be ill-defined (which reflects the singular nature
2386: of the wave equation (\ref{waveeq1})). Since $\phi$ is finite at
2387: $r=0$ we need $\del_{tt} \phi$ to be finite as well
2388:  \footnote{In the rest of this subsection ``finite'' should be
2389:  understood to mean ``finite at $r=0$.''}
2390: and thus we clearly need \be
2391:  L\,  \phi = {\rm finite}
2392:   \label{cond1} \ee
2393: where $\phi$ can be either of the initial conditions $\phi_i,\,
2394: \dot{\phi}_i$.
2395: 
2396: For RN from the leading behavior for $f_{\rm RN}$
2397: (\ref{fRN-lead}), we get that for a regular and generic $\phi$ the
2398: Laurent expansion of $L \phi$ starts at degree $(-4)$, and thus
2399: eq. (\ref{cond1}) represents 4 conditions (for
2400: degrees $-4 \le {\rm deg} \le -1$). Moreover, if we Taylor expand
2401: $\phi=\sum_{j=0}^\infty \phi^{(j)} \, r^j$ then at order ${\rm
2402: deg}$ the largest $j$ for which $\phi^{(j)}$ appears in the
2403: equation is $j={\rm deg}+6$, and so we have 4 (linear,
2404: homogeneous) conditions for the 6 coefficients $\phi^{(0)}, \dots,
2405: \phi^{(5)}$. These are actually the conditions found in the
2406: paragraph around eq. (\ref{Taylor-phi}), which limit our choice of
2407: initial conditions in the vicinity of $r=0$.
2408: 
2409: Similarly, for \Schw (\ref{cond1}) represents 3 conditions (for
2410: degrees $-3 \le {\rm deg} \le -1$) among the first 4 Taylor
2411: coefficients which agree with the findings in the paragraph after
2412: eq. (\ref{Taylor-phi}).
2413: 
2414: However, there are additional requirements: the constraint
2415: (\ref{cond1}) must be compatible with the time evolution according
2416: to the wave equation \be
2417:  L \phi = -\del_{tt} \phi \label{waveeq2} \ee
2418:  namely we must require
2419:  \be \del_{tt} [ L\, \phi] = {\rm finite}
2420:  \ee
2421:  By the wave equation (\ref{waveeq2}) this is equivalent to \be
2422:  L^2\,  \phi = {\rm finite}
2423:   \label{cond2} \ee
2424: This gives 4 additional conditions (in addition those of
2425: (\ref{cond1})) involving the first 12 Taylor coefficients for RN
2426: and 3 additional conditions for the first 8 Taylor coefficients
2427: for Schwarzschild.
2428: 
2429: \hspace{0.5cm}
2430: 
2431: Now we may readily generalize, and find \\
2432: {\bf Necessary boundary condition}:   \be L^n\, \phi= {\rm finite}
2433:  ~~ \forall n \ge 1 \label{bc} \ee
2434:  (at $r=0$).
2435: 
2436: \hspace{0.5cm}
2437: 
2438: This condition holds both for RN and Schwarzschild and is
2439: explicitly compatible with time evolution. Note that it can also
2440: be interpreted as $\del_{tt}^n\, \phi={\rm finite} ~~ \forall n
2441: \ge 1 $.  It is plausible to us that this condition is not only
2442: necessary but also sufficient for a unique time evolution as well
2443: since showing that $\phi$ has finite time derivatives of any order
2444: at the initial moment of time comes close to providing a
2445: well-defined time evolution, but we shall not attempt to pursue
2446: this point.
2447: 
2448: As we already mentioned, the boundary conditions (\ref{bc}) are
2449: automatically satisfied for characteristic b.c. (at past null
2450: infinity) and more generally b.c. which vanish at a neighborhood
2451: of $r=0$.
2452: 
2453: Another perspective is to consider a numerical computer
2454: implementation of this time evolution. One may either have no grid
2455: points at $r=0$ in which case the time evolution is well-defined,
2456: or put a grid point at $r=0$ and find its time increment as a
2457: limit of the time increments of neighboring points. The
2458: singularity will tend to ``expand'' numerical errors, but those
2459: should be possible to tame by decreasing the grid spacing. It
2460: would be interesting to analyze this further and/or to perform
2461: the implementation and determine the behavior at $r=0$.
2462: 
2463: 
2464: \subsection{General conditions for wave-regularity}
2465: 
2466: Having shown that \RN is wave-regular allowing for transmission
2467: across the singularity, we would like to abstract the general
2468: conditions on a spacetime to have such a resolution.
2469: 
2470: The crucial property of the RN singularity is that the two
2471: solutions $\phiom$ have a unique continuation across $r=0$, while
2472: {\it a priori} there could have been multi-valued functions such
2473: as the log's which appear in the Schwarzschild case. Hence a wave
2474: packet constructed from the $\phiom$ has a prospect of crossing
2475: the singularity in a well-defined manner. This is true not only in
2476: 4d RN where the time-separated ODE is regular but also for $d>4$
2477: RN where the ODE becomes regular-singular rather than smooth, but
2478: nevertheless the two characteristic exponents are integral and
2479: there are no $\log$ pieces so that the functions $\phiom$ are
2480: univalued in the vicinity of the singularity. The condition on the
2481: characteristic exponents can be generalized to allow for a
2482: reparameterization of coordinate  $r \to \tilde{r}\sim r^{c_1}$
2483: and a linear redefinition of the field $\phi \to r^{c_2} \,
2484: \phiom$ for some constants $c_1,\, c_2$. In general $\phiom$ can
2485: be series expanded as $\phiom = \tilde{r}^\rho\,
2486: \sum_{k=0}^{\infty}\, \phi^{(k)}\, \tilde{r}^{k\, \Delta}$ where
2487: $\rho=\rho_{1,2}$ is one of the two characteristic exponents and
2488: $\Delta$ is the ``series step-size'' ($\Delta=1$ when the
2489: functions in the differential equation are meromorphic) and
2490: $\phi^{(k)}$ are some constants. The quantity
2491: $(\rho_1-\rho_2)/\Delta$ is invariant under the double
2492: transformation above, and hence there exists a transformation such
2493: that all three $\rho_1,\, \rho_2$ and $\Delta$ are integral and
2494: the function is univalued exactly if the ratio above is rational.
2495: 
2496: \hspace{0.5cm}
2497: 
2498: We summarize the above by  \\
2499:  {\bf Necessary general condition}:
2500: the ``eigen-functions'' $\phiom$ should be univalued in the
2501: vicinity of the singularity. This is equivalent to: \begin{itemize}
2502:  \item The equation for $\phiom$ is either regular or regular-singular. If it
2503: is regular-singular we also require the next items:
2504:  \item The difference of
2505: characteristic exponents is commensurate with the ``series
2506: step-size'' (defined in the previous paragraph) \be
2507:  {\rho_2 - \rho_1 \over \Delta} \in \IQ~. \ee
2508: \item There are no $\log$ pieces in the solutions.
2509: \end{itemize}
2510: 
2511: \hspace{0.5cm}
2512: 
2513: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2514: \section{Summary and discussion}
2515: \label{discussion}
2516: 
2517: In this paper we saw that physical predictability can be restored
2518: in the neighborhood of the charged (RN) black-hole singularity as
2519: well as negative mass Schwarzschild, once one considers waves
2520: rather than particles. For RN this is done by gluing two
2521: spacetimes over the time-like singularity, thereby adding another
2522: asymptotic region. An observer at infinity in the additional
2523: region views a spacetime with a negative mass. Several
2524: alternatives exist for the total Penrose diagram depending on the
2525: relative size of $M^2$ and $Q^2$. This singularity appears to have
2526: the physical nature of a beam splitter.
2527: 
2528: For negative-mass Schwarzschild there is a natural ``regularity''
2529: boundary condition at the singularity cutting off spacetime there,
2530: and thus it can be physically interpreted as a perfectly
2531: reflecting mirror.
2532: 
2533: We subjected this picture to several tests by adding perturbations
2534: in section \ref{perturbations-section}. Considering a field with
2535: both mass and charge we found that such terms are subleading at
2536: the neighborhood of the singularity. Considering added
2537: interactions the field equation becomes non-linear and we were
2538: satisfied in confirming that there is no obstruction (such as
2539: creating a divergence) in extending the linear solutions to next
2540: order. Similarly we then considered the effect of non-linearities
2541: from back-reaction, and found that in some cases a simple argument
2542: protects the regularity of $\phi$ at the next order. Altogether
2543: the results were surprisingly resilient to perturbations.
2544: 
2545: It would be interesting to perform some additional tests and generalizations:
2546: 
2547: \begin{itemize}
2548: \item  Study fields with other spin.
2549: Of particular interest are the electromagnetic and gravitational
2550: fields, and also the Dirac $s=1/2$ field. A preliminary
2551: investigation of the electromagnetic and gravitational fields in
2552: d=4 indicates that their polar modes behave just as scalar fields,
2553: i.e. 75\% reflection and 25\% transmission for high energy at
2554: fixed $l$. The axial modes appear to be more subtle, however.
2555: 
2556: \item  Study other backgrounds such as extreme RN and rotating black-holes.
2557: \end{itemize}
2558: 
2559: Rotating black-holes deserve a special discussion here. They may
2560: be linked to our work in two different ways. First, the motivation
2561: to the present analysis partly emerges from the desire to explore
2562: the physical phenomena that may take place deep inside realistic
2563: rotating black holes. From this point of view the spherical
2564: charged black hole serves as a toy model for the more complicated,
2565: non-spherical, spinning black hole. In the second link we can
2566: employ the spinning black holes to test the singularity-resolution
2567: approach developed here. In the Kerr-Newman solution the $r=0$
2568: singularity forms a ring rather than a hypersurface. Once the
2569: field is decomposed into spheroidal-harmonic modes, the field
2570: equation for each mode is perfectly regular at $r=0$, hence there
2571: is no doubt about the proper continuation. Now, when the spin
2572: parameter $a$ is taken to vanish, the ring's radius shrinks to
2573: zero, and the spacetime becomes RN. One may therefore {\it define}
2574: the RN extension of the field beyond the $r=0$ singularity to be
2575: the limit $a \to 0$ of the corresponding field in the Kerr-Newman
2576: case. The obvious question is, therefore: does this procedure
2577: yield exactly the same extension as that constructed above? At
2578: least for a scalar field in 4d RN the answer is found to be
2579: positive. It still remains to check whether this is also the case
2580: in the $Q=0$ case. Namely, in the uncharged Kerr case, when $a \to
2581: 0$ and the spacetime becomes Schwarzschild, does one recover the
2582: full-reflection b.c. advocated above? This still needs to be
2583: verified.
2584: 
2585: 
2586: We would like to mention several other issues at the classical level:
2587: 
2588: \begin{itemize}
2589: \item  In all of our examples one of the spacetimes has a negative mass
2590: and a globally-naked singularity. Such physical objects raise
2591: problematic issues, and we name only a few: anti-gravity,
2592: acceleration reversed to force, and also inconsistency of their
2593: construction with the Cosmic Censorship conjecture (see
2594: ``Perspectives on $r<0$'' in section \ref{ReviewRN}).
2595: 
2596: It would be interesting to determine whether Nature allows the
2597: actual construction of any spacetime with a {\it wave-regular
2598: timelike singularity}. The mechanism that immediately suggests
2599: itself would be to reconsider a gravitational collapse of a
2600: charged spherical shell (in the positive-mass universe), leading
2601: to a RN black-hole geometry (outside the collapsing object).
2602: However, the instability of the inner horizon raises some doubts
2603: about whether the RN-like $r=0$ singularity will indeed form in
2604: this process.
2605: 
2606: \item  Implications for {\it Cosmic Censorship}. At the very least it
2607: shakes its rationale since given ``reasonable'' initial conditions
2608: Cosmic Censorship is supposed to ``protect us'' from loss of
2609: predictability, namely of losing unique time evolution, by
2610: forbidding naked singularities, but here we see that such
2611: singularities may not mean the loss of predictability after all.
2612: 
2613: \item  Since for RN we advocate a picture where there are two spacetimes
2614: which are consistently glued at the singularity
2615: --- and since a macroscopic measuring device cannot cross the singularity
2616: (at best it will bounce back, but it may also be destroyed by
2617: tidal forces, or be ``beam-splitted,'' in the worse case) one may
2618: take {\it two different points of view} on the physics
2619: corresponding to either of the two observers on the two sides of
2620: the singularity. Namely, the story of the resolution of each
2621: singularity will be told in two different versions corresponding
2622: to the two observers (this applies to both types of RN spacetime,
2623: namely $|Q|<|M|$ and $|Q|>|M|$ -- see figures
2624: \ref{PenroseRN},\ref{PenroseRN-ext}).
2625: 
2626: \end{itemize}
2627: 
2628: Going beyond classical GR there are open questions as to the quantum gravity
2629: consistency and properties:
2630: 
2631: \begin{itemize}
2632: \item We must note that our wave equation is outside its {\it domain of
2633: validity} near the singularity due to the presence of high
2634: curvature. However, a preliminary analysis shows that quantum
2635: corrections do not produce essential singularities in the
2636: equations and start altering the field only at a Planck (proper)
2637: distance from the singularity.
2638: 
2639: \item  Semi-classical quantization and Hawking evaporation (under study).
2640: 
2641: \item  As one passes to the quantum theory one may
2642: suspect that the pathologies associated with negative-mass spaces
2643: to only grow worse. Therefore, we need to be very cautious in
2644: discussing possible implications for {\it  ``legitimizing''
2645: spacetimes} with such singularities as possible solitons, namely
2646: which spacetimes should be considered to contribute to the path
2647: integral as admissible saddle points. Here we took an open-minded
2648: approach of exploration and we hope that the various interesting
2649: issues which get raised will be studied further.
2650: \end{itemize}
2651: 
2652: \vspace{0.5cm} \noindent {\bf Acknowledgements}
2653: 
2654: We would like to thank O. Aharony, J. Bekenstein, M. Berkooz, G.
2655: Gibbons, J. Katz, D. Kazhdan,  N. Itzhaki, D. Kutasov, M. Rozali
2656: and R.M. Wald for discussions. This work is supported in part by
2657: the Israeli Science Foundation. AG is supported in part by the
2658: Israel Academy of Sciences and Humanities -- Centers of Excellence
2659: Program, the German-Israel Bi-National Science Foundation, and the
2660: European RTN network HPRN-CT-2000-00122. BK is supported in part
2661: by the Israeli Science Foundation and by the Binational Science
2662: Foundation BSF-2002160. AS is supported in part by the Horowitz
2663: Foundation.
2664: 
2665: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2666: \appendix
2667: 
2668: \section{\KS coordinates}
2669: \label{KSapp}
2670: 
2671: For completeness, we recall here the derivation of the metrics in
2672: the \KS coordinates and the definition of the functions $g(r)$
2673: which are mentioned in the text and are used to implicitly define
2674: $r$ in these coordinates.
2675: 
2676: First one defines the ``tortoise'' coordinate $r^*$ by \be
2677:  dr^* := {dr \over f} ~. \ee
2678: $r^*$ diverges on horizons but is finite at $r=0$. Then one passes
2679: to light-cone coordinates \bea
2680:  v &:=& t+r^* \non
2681:  u &:=& t-r^*  \eea
2682: and finally in the neighborhood of a horizon with surface gravity
2683: $\kappa$ (for \RN we have $\kappa_\pm = (r_+-r_-)/(2\, r_\pm^2)$
2684: while for \Schw $\kappa=1/(2\, r_0)$)
2685:  the Kruskal-Szekeres coordinates are given by \bea
2686:  V &:=& \pm \exp (\pm \kappa\, v)  \non
2687:  U &:=& \pm \exp (\pm \kappa\, u)  \label{defUV} \eea
2688: and the sign inside the exponent is chosen such that the horizon
2689: is located at $UV=0$, namely it is X-shaped.
2690: 
2691: In these coordinates the metric is given by \be
2692:  ds^2 = -{f \over \kappa^2\, g}\, dU\, dV + r^2\, d\Omega^2 ~, \ee
2693:  where the function $g(r)$ is defined by \be
2694:  g(r) := \exp (\pm 2\, \kappa\, r^* ) \ee
2695:  and the sign is chosen to be the same as that for $V$ in (\ref{defUV}).
2696: $r$ is not
2697: a coordinate anymore, but rather it is implicitly defined in terms
2698: of the $U,\, V$ coordinates by \be
2699:  -U\, V = g(r) ~. \ee
2700: Note that by construction $g(r)$ has a zero at the horizon and
2701: thus the horizon is explicitly smooth in these coordinates since
2702: in the prefactor of $dU\, dV$ the zero in $f$ gets cancelled
2703: against the zero in $g$.
2704: 
2705: In \RN there are two horizons at $r_\pm$, and accordingly two
2706: Kruskal-Szekeres-like planes for $(U_\pm,\, V_\pm)$, and two
2707: functions $g_\pm (r)$. The ``upper plus'' quadrant ($U_+,V_+>0$)
2708: is glued to the ``lower minus'' quadrant ($U_-,V_-<0$) according
2709: to \bea
2710:  V_+^{1/\kappa_+} &=& (-V_- )^{1/\kappa_-} \non
2711:  U_+^{1/\kappa_+} &=& (-U_- )^{1/\kappa_-}  \eea
2712: and similarly one may continue and glue the upper minus quadrant
2713: to another copy of the plus plane, creating the maximal analytic
2714: extension of \RN which consists of an infinite chain of
2715: alternating plus and minus planes. $g_\pm(r)$ are given by (see
2716: figure \ref{gRNfigure}) \bea
2717:  g_+(r) &=& \({r \over r_+}-1 \) \, \({r \over r_-}-1 \)^{-{\kappa_+ \over
2718: \kappa_-}} \, \exp (2\, \kappa_+\, r) \non
2719:  g_-(r) &=& \(1-{r \over r_-}\) \, \(1-{r \over r_+}\)^{-{\kappa_- \over
2720: \kappa_+}} \, \exp
2721:  (-2\, \kappa_-\, r) \eea
2722: While for Schwarzschild we have (see figure \ref{gSchw-figure})
2723: \be
2724:  g_{\rm schw}(r) = \({r \over r_0}-1 \)\, \exp(r/r_0) ~. \ee
2725: Finally, in order to get the Penrose diagram (see figures
2726: \ref{PenroseRN},\ref{PenroseSchw},\ref{PenroseRN-ext}) one
2727: customarily takes the conformal transformation \bea
2728:  U_P = {\rm tg}^{-1} (U) \non
2729:  V_P = {\rm tg}^{-1} (V) \eea
2730: although any other transformation which maps the real line to an
2731: interval is admissible.
2732: 
2733: 
2734: \section{Regular singularities of ordinary differential equations}
2735: \label{reg-sing-app}
2736: 
2737: Let us briefly recall the definitions.
2738: A linear second order differential equation \be
2739:  [a(r)\, \del_{rr} + b(r)\, \del_r + c(r)]\, \phi=0 \ee
2740: is regular-singular at a point $r_0$, which we will assume without
2741: loss of generality to be $r_0=0$, if after normalization such
2742:  that $a(0)=1$ there are singularities at $r=0$ in $b(r)$ or
2743: $c(r)$ of limited type: $b(r)$ may have at most a first order
2744: pole, and $c(r)$ at most a second order pole. This guarantees that
2745: the solutions will have at most poles or branch cuts at $r=0$ but
2746: not an essential singularity. More specifically the two solutions
2747: have the leading behavior \be
2748:  \phi_{1,2} \sim r^{\rho_{1,2}} ~, \ee
2749:  where $\rho_{1,2}$ are called the characteristic exponents, and they are
2750: the solutions to the quadratic equation \be
2751:  a(r_0) \rho\, (\rho-1) + b(r_0)\, r  \rho + c(r_0)\, r^2  =0 \ee
2752:  gotten from substituting the leading
2753: behavior above into the equation. When $\rho_1=\rho_2=\rho$ the
2754: leading behavior is $\phi_1=r^\rho, ~ \phi_2=r^\rho \, \log(r)$.
2755: Finally, one can expand the solutions into a series $\phi = r^\rho
2756: \sum_{j=0}^{\infty}\, \phi^{(j)}\, r^j$ where $\phi^{(j)}$ are
2757: some constants, except that in the case when $\rho_2-\rho_1$ is a
2758: positive integer $\phi_1$ may contain also a piece proportional to
2759: $\log(r)\, \phi_2$ and thus contains a log.
2760: 
2761: \section{Basis functions for the characteristic formulation}
2762: \label{char-appendix}
2763: 
2764: Here we define the basis functions $\phi _{(\omega _{v},0)}^{*},\,
2765: \phi _{(0,\omega _{u})}^{*}$ for the characteristic formulation
2766: (see subsection \ref{char-subsection}). Since the radial equation
2767: is a second-order ODE, the radial functions $\phi _{\omega
2768: }^{*}(r^*)$ for a given $\omega $ form a two-parameter family.
2769: Owing to the asymptotic behavior of the radial functions, Eq.
2770: (\ref{gen-asymptotic}), we may choose two basis functions $\phi
2771: _{\omega (1,0)}^{*}(r^*)$ and $\phi _{\omega (0,1)}^{*}(r^*)$,
2772: defined as follows: $\phi _{\omega (1,0)}^{*}(r^*)$ is the radial
2773: function which at $r^*\to +\infty $ (the inner horizon) has the
2774: asymptotic form
2775: \[
2776: \phi _{\omega (1,0)}^{*}\cong T(\omega ){\rm exp}(-i\,\omega
2777: \,r^*)
2778: \]
2779: and at $r^*\to -\infty $ (the negative-$r$ asymptotically-flat
2780: region) has the asymptotic form
2781: \[
2782: \phi _{\omega (1,0)}^{*}\cong {\rm exp}(-i\,\omega \,r^*)+R(\omega
2783: ){\rm exp} (+i\,\omega \,r^*),
2784: \]
2785: where $T(\omega )$ and $R(\omega )$ are two unconstrained
2786: coefficients (these coefficients turn out to be the transmission
2787: and reflection coefficient; see section
2788: \ref{cross-section-section}). Similarly, $\phi_{\omega
2789: (0,1)}^{*}(r^*)$ is the radial function which at $r^* \to -\infty
2790: $ has the asymptotic form
2791: \[
2792: \phi _{\omega (0,1)}^{*}\cong T^{\prime }(\omega ){\rm
2793: exp}(+i\,\omega \,r^*)
2794: \]
2795: and at $r^*\to +\infty $ has the asymptotic form
2796: \[
2797: \phi _{\omega (0,1)}^{*}\cong {\rm exp}(+i\,\omega
2798: \,r^*)+R^{\prime }(\omega ) {\rm exp}(-i\,\omega \,r^*),
2799: \]
2800: where again $T^{\prime }(\omega )$ and $R^{\prime }(\omega )$ are
2801: two unconstrained coefficients. Now, the two desired solutions
2802: $\phi _{(\omega _{v},0)}^{*}$ and $\phi _{(0,\omega _{u})}^{*}$
2803: are simply given by
2804: \begin{eqnarray*}
2805: \phi _{(\omega _{v},0)}^{*}(r^*,t) &=&\phi _{\omega
2806: _{v}(1,0)}^{*}(r^*){\rm exp}(i\,\omega _{v}\,t)\,,\, \\
2807: \phi _{(0,\omega _{u})}^{*}(r^*,t) &=&\phi _{\omega
2808: _{u}(0,1)}^{*}(r^*){\rm exp}(i\,\omega _{u}\,t).
2809: \end{eqnarray*}
2810: As one can easily verify, these two solutions take the asymptotic
2811: forms
2812: \[
2813: \phi _{(\omega _{v},0)}^{*}\cong T(\omega _{v}){\rm exp}(i\,\omega
2814: _{v}\,v)\qquad \qquad (r^*\to +\infty ),
2815: \]
2816: \[
2817: \phi _{(\omega _{v},0)}^{*}\cong {\rm exp}(i\,\omega
2818: _{v}\,v)+R(\omega ){\rm exp}(i\,\omega _{v}\,u)\qquad \qquad
2819: (r^*\to -\infty ),
2820: \]
2821: and
2822: \[
2823: \phi _{(0,\omega _{u})}^{*}\cong T^{\prime }(\omega _{u}){\rm
2824: exp}(i\,\omega _{u}\,u)\qquad \qquad (r^*\to -\infty ),
2825: \]
2826: \[
2827: \phi _{(0,\omega _{u})}^{*}\cong {\rm exp}(i\,\omega
2828: _{u}\,u)+R^{\prime }(\omega ){\rm exp}(i\,\omega _{u}\,v)\qquad
2829: \qquad (r^*\to +\infty ),
2830: \]
2831: in agreement with the above definitions of $\phi _{(\omega
2832: _{v},0)}^{*}$ and $\phi _{(0,\omega _{u})}^{*}$.
2833: 
2834: 
2835: \section{Bessel functions}
2836: \label{BesselApp}
2837: 
2838: Let us assemble a few useful properties of the Bessel functions. A
2839: Bessel function of order $n$ satisfies the equation
2840: $[\del_{xx}+(1/x)\, \del_x + 1-n^2/x^2 ]\,J_n\, =0$. It will be
2841: more convenient for us to use the equivalent definition \be
2842:   \br \del{xx} + \(1- {n^2-1/4 \over x^2} \)  \kt \, \sqrt{x} J_n(x)
2843:  =0 ~. \ee
2844: For small $x$ \be J_n(x) = x^n\, A^{J}_n(x^2) ~,
2845: \label{BesselSing} \ee
2846:  where $A^{J}_n(x^2)$ are certain functions of $x^2$ analytic at $x^2=0$.
2847: The asymptotic behavior is better described by \be
2848:  Y_n=(J_n\, \cos (n\, \pi) - J_{-n})/\sin(n\, \pi) \ee
2849:  and one has \be
2850:  J_n \pm i\, Y_n \simeq \sqrt{2/(\pi\, x)}~ \exp \left[ \pm i(x-n\,
2851:  \pi/2-\pi/4) \right]
2852:  \label{BesselAsymp}\ee
2853:  as $x \to +\infty$.
2854: 
2855: 
2856: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2857: \begin{thebibliography}{99}
2858: 
2859: \bibitem{Polchinski}
2860: J.~Polchinski, ``String Theory. Vol. 2: Superstring Theory And
2861: Beyond,'' ``String Theory. Vol. 1: An Introduction To The Bosonic
2862: String.''
2863: 
2864: \bibitem{GRS-review}
2865: A.~Giveon, E.~Rabinovici and A.~Sever, ``Strings in singular
2866: time-dependent backgrounds,'' Fortsch.\ Phys.\  {\bf 51}, 805
2867: (2003) [arXiv:hep-th/0305137].
2868: %%CITATION = HEP-TH 0305137;%%
2869: 
2870: \bibitem{dvv}
2871: R.~Dijkgraaf, H.~Verlinde and E.~Verlinde, ``String propagation in
2872: a black hole geometry,'' Nucl.\ Phys.\ B {\bf 371} (1992) 269.
2873: 
2874: \bibitem{GiveonRabinoviciSever}
2875: A.~Giveon, E.~Rabinovici and A.~Sever, ``Beyond the singularity of
2876: the 2-D charged black hole,'' JHEP {\bf 0307}, 055 (2003)
2877: [arXiv:hep-th/0305140].
2878: %%CITATION = HEP-TH 0305140;%%
2879: 
2880: \bibitem{bardeen}  J. M. Bardeen, Nature {\bf 226} 64 (1970).
2881: 
2882: \bibitem{thorne}  K. S. Thorne, Astrophys. J. {\bf 191} 507 (1974).
2883: 
2884: \bibitem{central}  For a recent observation see R. Genzel, R. Schoedel, T.
2885: Ott, A. Eckart, T. Alexander, F. Lacombe, D. Rouan, B. Aschenbach,
2886: Nature {\bf 425}, 934 (2003).
2887: 
2888: \bibitem{IshibashiHosoya}
2889: A.~Ishibashi and A.~Hosoya, ``Who's afraid of naked singularities?
2890: Probing timelike singularities  with finite energy waves,'' Phys.\
2891: Rev.\ D {\bf 60}, 104028 (1999) [arXiv:gr-qc/9907009].
2892: %%CITATION = GR-QC 9907009;%%
2893: 
2894: \bibitem{Wald1980}
2895: R.~M.~Wald, ``Dynamics in nonglobally hyperbolic, static
2896: space-times,'' J.\ Math.\ Phys.\  {\bf 21}, 2802 (1980).
2897: %%CITATION = JMAPA,21,2802;%%
2898: 
2899: \bibitem{HorowitzMarolf}
2900: G.~T.~Horowitz and D.~Marolf, ``Quantum probes of space-time
2901: singularities,'' Phys.\ Rev.\ D {\bf 52}, 5670 (1995)
2902: [arXiv:gr-qc/9504028].
2903: %%CITATION = GR-QC 9504028;%%
2904: 
2905: %\cite{Peeters:1994jz}
2906: \bibitem{Peeters:1994jz}
2907: K.~Peeters, C.~Schweigert and J.~W.~van Holten, ``Extended
2908: geometry of black holes,'' Class.\ Quant.\ Grav.\  {\bf 12}, 173
2909: (1995) [arXiv:gr-qc/9407006].
2910: %%CITATION = GR-QC 9407006;%%
2911: 
2912: \bibitem{Jacobson}
2913: T.~Jacobson, ``Semiclassical decay of near-extremal black holes,''
2914: Phys.\ Rev.\ D {\bf 57}, 4890 (1998) [arXiv:hep-th/9705017].
2915: %%CITATION = HEP-TH 9705017;%%
2916: 
2917: \bibitem{Lynden-BellKatz}
2918:  D.~Lynden-Bell and J.~Katz,
2919:  ``Geometric extension through Schwarzschild r = 0,''
2920: Mon.\ Not.\ Roy.\ Astron.\ Soc.\ (1990) { \bf 247}, 651.
2921: 
2922: \bibitem{Lynden-BellKatz2}
2923:  D.~Lynden-Bell, J.~Katz and C. Hellaby,
2924:  ``Correction to geometric extension through Schwarzschild r = 0,''
2925:   Mon.\ Not.\ Roy.\ Astron.\ Soc.\ (1993)
2926: {\bf 262}, 325.
2927: 
2928: \bibitem{Reissner} H.~Reissner, ``\"{U}ber die Eigengravitation
2929: des elektrischen Felds nach den Einsteinschen Theorie,'' Ann.\
2930: Phys.\, {bf 50}, 106-120 (1916).
2931: 
2932: \bibitem{Nordstrom} G.~Nordstr\"{o}m, ``On the energy of the
2933: gravitational field in Einstein's theory,'' Proc.\ Kon.\ Ned.\
2934: Akad.\ Wet.\, {\bf 20}, 1238-1245 (1918).
2935: 
2936: \bibitem{Schw} K.~Schwarzschild, ``Uber das
2937: Gravitationsfeld eines Massenpunktes nach der Einsteinschen
2938: Theorie,'' Sitzber.\ Deut.\ Akad.\ Wiss.\ Berlin, Kl.\
2939: Math.\--Phys.\ Tech.\, 189-196 (1916). \\
2940: English translation: Gen.\ Relativ.\ Gravit.\ {\bf 35}, 951 (2003)
2941: [physics/9905030].
2942: 
2943: \bibitem{Kruskal}
2944: M.~D.~Kruskal, ``Maximal Extension Of Schwarzschild Metric,''
2945: Phys.\ Rev.\  {\bf 119}, 1743 (1960).
2946: %%CITATION = PHRVA,119,1743;%%
2947: 
2948: \bibitem{Szekeres} G.~Szekeres, ``On the singularities of a
2949: Riemannian manifold,'' Publ.\ Mat.\ Debrecen\, {\bf 7}, 285-301,
2950: (1960).
2951: 
2952: \bibitem{Israel}
2953: W.~Israel, ``Singular Hypersurfaces And Thin Shells In General
2954: Relativity,'' Nuovo Cim.\ B {\bf 44S10}, 1 (1966) [Erratum-ibid.\
2955: B {\bf 48}, 463 (1967\ NUCIA,B44,1.1966)].
2956: %%CITATION = NUCIA,B44S10,1;%%
2957: 
2958: \bibitem{Visser}
2959: M.~Visser, ``Traversable Wormholes From Surgically Modified \\
2960: Schwarzschild Space-Times,'' Nucl.\ Phys.\ B {\bf 328}, 203
2961: (1989).
2962: %%CITATION = NUPHA,B328,203;%%
2963: 
2964: \bibitem{negative-supergravity}  T. Hertog, G. T. Horowitz, and K. Maeda,
2965: ``Negative Energy in String Theory and Cosmic Censorship
2966: Violation'', hep-th/0310054.
2967: 
2968: \bibitem{Penrose}  R. Penrose, in Battelle Rencontres, 1967 lectures in
2969: mathematics and physics , edited by C. M. DeWitt and J. A. Wheeler
2970: (Benjamin, New York, 1968), P. 222 .
2971: 
2972: \bibitem{PI}  E.~Poisson and W.~Israel,
2973: ``Internal Structure Of Black Holes,'' Phys.\ Rev.\ D {\bf 41},
2974: 1796 (1990)
2975: %%CITATION = PHRVA,D41,1796;%%
2976: and references therein.
2977: 
2978: \bibitem{weak-singularity}  A. Ori, ``Inner structure of a
2979: charged black hole: An exact mass-inflation solution,''
2980: Phys. Rev. Lett. 67 , 789 (1991).
2981: 
2982: \bibitem{burko}
2983: L.~M.~Burko, ``Structure of the black hole's Cauchy horizon
2984: singularity,'' Phys.\ Rev.\ Lett.\  {\bf 79}, 4958 (1997)
2985: [arXiv:gr-qc/9710112].
2986: %%CITATION = GR-QC 9710112;%%
2987: 
2988: \bibitem{spinning}  This is also the situation in the more realistic,
2989: spinning black-hole case: see e.g.
2990: A. Ori, ``Structure of the singularity inside a
2991: realistic rotating black hole,'' Phys. Rev. Lett. 68 ,
2992: 2117 (1992); A.~Ori,
2993:  ``Evolution Of Linear Gravitational And Electromagnetic
2994:  Perturbations Inside A
2995: Kerr Black Hole,'' Phys.\ Rev.\ D {\bf 61}, 024001 (2000).
2996: %%CITATION = PHRVA,D61,024001;%%
2997: 
2998: \bibitem{LL35}
2999: L.~D.~Landau and E.~M.~lifshitz, ``Quantum mechanics,'' {\it
3000: Pergamon} (1977), \S35.
3001: 
3002: \bibitem{MyersPerry}
3003: R.~C.~Myers and M.~J.~Perry, ``Black holes in higher dimensional
3004: space-times,'' Annals Phys.\  {\bf 172}, 304 (1986).
3005: %%CITATION = APNYA,172,304;%%
3006: 
3007: 
3008: \bibitem{nappiyost}
3009: M.~D.~McGuigan, C.~R.~Nappi and S.~A.~Yost, ``Charged black holes
3010: in two-dimensional string theory,'' Nucl.\ Phys.\ B {\bf 375}, 421
3011: (1992) [arXiv:hep-th/9111038].
3012: %%CITATION = HEP-TH 9111038;%%
3013: 
3014: \bibitem{bars}
3015: I.~Bars and K.~Sfetsos, ``Conformally exact metric and dilaton in
3016: string theory on curved space-time,'' Phys.\ Rev.\ D {\bf 46}
3017: (1992) 4510 [arXiv:hep-th/9206006].
3018: %%CITATION = HEP-TH 9206006;%%
3019: 
3020: \bibitem{giveon}
3021: A.~Giveon, ``Target space duality and stringy black holes,'' Mod.\
3022: Phys.\ Lett.\ A {\bf 6}, 2843 (1991).
3023: 
3024: \bibitem{gpr}
3025: A.~Giveon, M.~Porrati and E.~Rabinovici, ``Target space duality in
3026: string theory,'' Phys.\ Rept.\  {\bf 244}, 77 (1994)
3027: [arXiv:hep-th/9401139].
3028: 
3029: 
3030: \bibitem{ReedSimon}
3031:  M.~Reed and B.~Simon,
3032:  ``Methods of modern mathematical physics, v2: Fourier analysis
3033:  and self-adjointness'', {\it Academic Press} (1975).
3034: 
3035: \end{thebibliography}
3036: 
3037: \end{document}
3038: