1: \documentclass[11pt]{article}
2: \usepackage{amsfonts,amssymb,latexsym}
3: \usepackage{amsmath,amsgen,graphicx}
4: %\usepackage{theorem,showkeys}
5: \addtolength{\oddsidemargin}{-.687in}
6: \addtolength{\textwidth}{1.25in}
7: \addtolength{\topmargin}{-.75in}
8: \addtolength{\textheight}{1.25in}
9: \newtheorem{claim}{Claim}{\it}{\rm}
10: \newtheorem{lemma}{LEMMA}[section]{\bf}{\it}
11: \newtheorem{prop}{PROPOSITION}{\bf}{\it}
12: \newtheorem{remark}{Remark}{\bf}{\rm}
13: \newtheorem{theorem}{THEOREM}{\bf}{\it}
14: \newcommand{\sphlap}{\displaystyle{\not\!\!\Delta}}
15: \newcommand{\sphgrad}{\displaystyle{\not\!\nabla}}
16: \newcommand{\ds}{\displaystyle}
17: \newcommand{\nab}{\nabla}
18: \newcommand{\half}{\frac{1}{2}}
19: \newcommand{\p}{\partial}
20: \newcommand{\A}{{\cal A}}
21: \newcommand{\E}{{\cal E}}
22: \newcommand{\Hi}{\mathbf{H}}
23: \newcommand{\Hc}{\mathcal{H}}
24: \newcommand{\M}{{\cal M}}
25: \newcommand{\RR}{{\mathbb R}}
26: \newcommand{\ZZ}{{\mathbb Z}}
27: \newcommand{\BS}{{\mathbb S}}
28: \newcommand{\mb}[1]{\mathbf{#1}}
29: \newcommand{\rad}{\mathrm{rad}}
30: \newcommand{\qed}{\par\hfill\rule{2mm}{2mm}}
31: \begin{document}
32:
33: \title{
34: Scalar waves on a\\ naked-singularity background}
35:
36: \author{
37: John G. Stalker$^\mathbf{a}$
38: \and
39: A. Shadi Tahvildar-Zadeh\thanks{Research supported in part by the National Science Foundation grant DMS 0301207.}$^{\ ,\mathbf{b}}$
40: }
41:
42: \date{March 2004}
43: \maketitle
44: \begin{abstract}
45:
46: We obtain global space-time weighted-$L^2$ (Morawetz) and $L^4$ (Strichartz) estimates for a massless chargeless spherically symmetric scalar field propagating on a super-extremal (overcharged)
47: Reissner-Nordstr\"om background. To do this we first discuss the well-posedness of the Cauchy problem for scalar fields on non-globally hyperbolic manifolds, review the role played by the Friedrichs extension, and go over the construction of the function spaces involved. We then show how to transform this problem to one about the wave equation on the Minkowski space with a singular potential, and prove that the potential thus obtained satisfies the various conditions needed in order for the estimates to hold.
48: \end{abstract}
49:
50:
51:
52: \section{Introduction}\label{sec:intro}
53:
54: There have been many studies of the well-posedness and decay of scalar fields in a given
55: space-time whose metric satisfies the Einstein equations of general relativity,
56: both as a precursor to the study of the stability of that metric,
57: and as a means of probing various censorship conjectures. Although our main goal in this paper
58: is to obtain estimates for scalar fields, we need to address the issue
59: of well-posedness as well.
60:
61: \subsection{Well-posedness of the Cauchy problem for scalar fields}
62:
63: This question arises in studying the phenomenon of singularities in general
64: relativity. The classical notion of a singularity of space-time is that of geodesic
65: incompleteness. The weak cosmic
66: censorship conjecture states that generically, singularities of spacetime must be hidden inside
67: black holes, instead of being ``naked'', i.e. visible to distant observers. The strong form of
68: this conjecture posits that generically, spacetimes must be globally hyperbolic, i.e. possess a
69: complete
70: spacelike hypersurface such that every causal curve in the manifold intersects it at exactly
71: one point. Global hyperbolicity ensures that the spacetime has deterministic dynamics, since
72: there is a Cauchy surface whose domain of dependence is the entire spacetime. In the case
73: where the space-time is not globally hyperbolic, which can happen for example if the metric
74: is singular\footnote{The minimum regularity required of the metric for
75: the local existence and uniqueness of geodesics
76: to hold is $C^{1,\alpha}$, $\alpha>0$, so that any point in the manifold where the metric fails to
77: be $C^{1,\alpha}$
78: needs to be removed, which can easily result in the loss of global hyperbolicity \cite{Cla98}. We note however, that the singularities of metrics considered in this paper are in fact much stronger, and the curve-integrability conditions of \cite{Cla98} do not hold for these examples.} on a
79: time-like curve, well-posedness of the Cauchy problem for the scalar wave
80: equation has been suggested
81: \cite{Wal80} as a substitute for geodesic completeness in determining how singular the
82: spacetime actually is, and perhaps to see if quantum effects have any chance of
83: regularizing the dynamics and restoring predictability \cite{HorMar95}. This is
84: particularly relevant if censorship conjectures somehow fail to be true and naked
85: singularities turn out to be more abundant than otherwise allowed by them.
86:
87: Let us now recall the set-up from \cite{Wal80}: Let $(\M,g)$ be a
88: static, stably causal space-time, i.e., one that admits a hypersurface-orthogonal Killing vector
89: field $T^\mu$ whose
90: orbits are complete and everywhere timelike. A massless scalar field on $\M$ satisfies the
91: equation
92: \begin{equation}\label{eq:ur}
93: g^{\mu\nu}\nab_\nu \nab_\mu \psi = 0.
94: \end{equation}
95: Suppose we specify initial data for $\psi$ on a hypersurface $\Sigma$ that is everywhere
96: orthogonal to $T^\mu$. If $\M$ is not globally hyperbolic, $\Sigma$ will not be a Cauchy
97: surface and data on $\Sigma$ will determine $\psi$ only on the domain of dependence
98: $D(\Sigma)$. The aim of \cite{Wal80} was to define a physically sensible recipe for
99: determining $\psi$ everywhere in $\M$. To this end, one
100: rewrites
101: (\ref{eq:ur}) in the form
102: \[
103: \partial_t^2 \psi = \alpha D^a(\alpha D_a\psi)
104: \]
105: Here $\alpha = \sqrt{-g(T,T)}$, $t$ is the Killing parameter, and $D_a$ is the covariant
106: derivative of the Riemannian metric induced on $\Sigma$. One then views
107: $$A:= -\alpha D^a(\alpha D_a)$$ as an operator on the Hilbert space
108: $$\Hc = L^2(\Sigma,\alpha^{-1}d\sigma),$$ where
109: $d\sigma$ denotes the induced volume form of $\Sigma$. The point of this definition is that with respect
110: to the inner product of $\Hc$, $A$ will now be symmetric and positive, although not
111: yet self-adjoint since the initial domain of $A$ would have to consist of sufficiently
112: smooth functions. If we take this initial domain to be $C^\infty_c(\Sigma)$
113: (smooth functions of compact support on $\Sigma$) then $A$ is also densely defined,
114: and the equation we want to solve is
115: \begin{equation}\label{eq:Apsi}
116: (\partial_t^2 +A) \psi = 0,\qquad \psi_{\left|_\Sigma\right.} = f,
117: \qquad \partial_t\psi_{\left|_\Sigma\right.} = g.
118: \end{equation}
119: It is then a classical result \cite{ReeSimII}, that the above problem is well-posed,
120: provided one replaces $A$ in this
121: equation with (one of) its self-adjoint extension(s) $A_E$. Such self-adjoint
122: extensions are guaranteed to exist for real, symmetric operators \cite{Neu30}.
123: The extension may not be unique though, which would imply that there is still an
124: ambiguity about the dynamics. This may be interpreted as having to specify
125: boundary conditions for the scalar field ``on the singularity''. One finds
126: that there are three proposals for removing the ambiguity and restoring determinism
127: to the dynamics of scalar fields:
128:
129: {\bf (1)} One can of course restrict attention only to the cases where the
130: self-adjoint extension is unique (the so-called essentially self-adjoint case), and
131: declare that spacetimes where the operator $A$ is not essentially self-adjoint
132: are ``quantum-mechanically singular'' \cite{HorMar95} (meaning the singularity
133: remains even if quantum particle dynamics, i.e waves, are considered in place of
134: classical particle dynamics, i.e. geodesics). In this approach, the naked singularities
135: present at the center in both the negative mass Schwarzschild ($m<0$) and the
136: super-extremal Reissner-Nordstr\"om ($|e|>m$) spacetimes are quantum-mechanically
137: singular, and nothing more can be said about the evolution of scalar fields on these
138: backgrounds.
139:
140: {\bf (2)} Another possibility is to characterize the singularity that leads to
141: non-unique self-adjoint extensions for $A$ as having a $U(N)$ ``hair,'' \cite{IshHos99}
142: where $N$
143: is the common value of the two deficiency indices $n_\pm := \dim(\ker(A^*\pm i))$ of $A$. This
144: is because on the one hand there is a one-to-one correspondence between the self-adjoint extensions of $A$
145: and unitary maps from $\ker(A^*+i)$ onto $\ker(A^*-i)$ \cite{Neu30}, and
146: on the other hand, as
147: shown in \cite{IshWal03}, any ``reasonable'' way of defining the dynamics on the whole
148: spacetime, would necessarily have to arise from {\em some} self-adjoint extension of $A$.
149: We will show that the naked singularity of the super-extremal Reissner-Nordstr\"om solution
150: is in this sense, quite hairy (see Remark~\ref{rem:hair}).
151:
152:
153: {\bf (3)} A third approach is to distinguish one self-adjoint extension from among all the
154: possible ones, as being somehow more ``natural'' or ``physical''. It was suggested in
155: \cite{Wal80} that the Friedrichs extension of $A$, i.e. the one coming from extending
156: the corresponding quadratic form, is such a natural choice. We note that the Friedrichs extension of $A$ always exists, and is unique by construction\footnote{The claimed non-uniqueness of the $H^1$-based extension (which is equivalent to the Friedrichs extension) of this operator in the case of superextremal Reissner-Nordstr\"om, \cite[\S IV.A.2]{IshHos99}, is due to the authors' error in not realizing the Sobolev space $H^1$ as the closure of $C^\infty_c$ functions under the corresponding norm (see Section 3).}. Moreover, as was shown in \cite{Seg03}, the Friedrichs extension is
157: the only self-adjoint extension of $A$ under which the resulting dynamics of (\ref{eq:Apsi}) agrees with that
158: of the corresponding first-order (Hamiltonian) formulation,
159: \[
160: \partial_t \Psi = -h \Psi,
161: \]
162: with
163: \[
164: \Psi = \left(\begin{array}{c}\psi\\ \alpha^{-1}\partial_t \psi\end{array}\right),
165: \qquad h = \left(\begin{array}{cc} 0 & -\alpha \\ \alpha^{-1}A & 0\end{array}\right),
166: \]
167: obtained via the (unique) skew-adjoint extension of the corresponding operator $h$.
168:
169: In this paper we will take the third approach, and work with the
170: Friedrichs extension of $A$. This is mainly because finiteness of the $H^1$
171: norm is used several times in the course of the proof of our estimates,
172: and among the extensions of $A$, the Friedrichs extension is the only
173: one whose domain is contained in $H^1$. It may be that some of our
174: estimates hold for the other extensions of $A$ as well, but our proof
175: does not extend to those cases.
176:
177: \subsection{Estimates for scalar fields}
178:
179: The next natural question to consider for scalar waves after well-posedness is
180: obtaining estimates for them.
181: This is relevant among other things to the question of stability of the metric as a
182: solution of the
183: field equations, since expanding the perturbation functions in tensor harmonics yields
184: a sequence of scalar wave equations \cite{Mon74}, and having good estimates for each
185: one of them seems necessary--although not sufficient--for proving linear
186: stability of the metric under a small perturbation of its data. Moreover, the question
187: of stability has an obvious connection to cosmic censorship if the solution
188: to be perturbed happens to have a naked singularity.
189:
190: Most of the results obtained so far in the direction of estimates
191: concern either boundedness of the field \cite{KayWal87,Whi89,
192: Bey01} or its dispersion, i.e. the decay rate, with respect to the foliation parameter, of the
193: supremum of the field on the slices of a given time-like or null foliation of
194: a portion of the space-time
195: \cite{Pri72,Bic78,KoyTom01a,KoyTom01b,FKSY02,CYDL03,Poi03,MacSta04,DafRod03}.
196: In this paper we take a different approach to this problem, by
197: obtaining estimates that bound the {\em global} space-time $L^4$-norm (and weighted-$L^2$
198: norm) of the field, in terms
199: of appropriate norms of its data on a given Cauchy hypersurface (see \cite{BluSof03,BluSof03b} for
200: a similar approach). Such Strichartz (resp. Morawetz) estimates,
201: as they are known in the literature,
202: are a staple of the modern theory of partial differential equations, and have proved
203: very useful in the analysis of nonlinear evolution problems. We will be looking at
204: the simplest non-trivial situation possible, namely, a spherically symmetric scalar field
205: on a given spherically symmetric and static background space-time\footnote{It is possible to
206: extend our result to fields that are a finite sum of spherical harmonics, as in \cite{FKSY02}, but we
207: do not pursue it here since this approach is not likely to yield an estimate for the general,
208: nonsymmetric case.}.
209: Furthermore, the background here is allowed
210: to have a naked singularity, the main example we have in mind being the super-extremal
211: (naked) Reissner-Nordstr\"om solution of the Einstein-Maxwell system. Our main result
212: is
213: \begin{theorem}\label{thm:intro}
214: Let $(\M,g)$ be the Reissner-Nordstr\"om space-time manifold, with mass $m$ and charge $e$,
215: such that
216: \[
217: |e|> 2m,
218: \]
219: and let $\psi$ be a massless, chargeless, spherically symmetric,
220: scalar field on $\M$. Then there exists a constant $C>0$
221: (which may
222: depend on $e$ and $m$) such that
223: \begin{equation}\label{est:intro}
224: \| \psi\|_{L^4(d\mu_g)} + \|\rho^{-1}\psi\|_{L^2(d\mu_g)} \leq C
225: \mb{E}_{1/2}(\psi),
226: \end{equation}
227: where $\rho$ denotes the area-radius coordinate on $\M$, defined in Section~2 and
228: $\mb{E}_{1/2}(\psi)$ is the
229: conserved $\half$-energy of the field, defined in Section~\ref{sec:energy}.
230: \end{theorem}
231: The outline of this paper is as follows: In Section 2 we introduce the family of static
232: spherically-symmetric Lorentzian manifolds, define a global system of coordinates on them,
233: and write down the evolution equation satisfied by a scalar field on such a manifold.
234: In Section 3 we introduce the notion of self-adjoint extension that is necessary to make
235: that evolution problem well-posed, and define the function spaces that are going to be
236: used. In Section 4 we state the various conditions that need to be satisfied by the
237: metric coefficients of a manifold in the family we are considering, such that a scalar
238: field on it will satisfy the estimate (\ref{est:intro}). This is then proved by
239: transforming the problem to one about the flat wave equation with a potential,
240: appealing to
241: the result in \cite{BPST2}, and transforming back to the original problem. The
242: last section contains the proof of the fact that the metric coefficients of the
243: super-extremal Reissner-Nordstr\"om
244: solution satisfy the above mentioned conditions. This is accomplished by reformulating
245: the problem in the language of real algebraic curves and appealing to the compactness result
246: of~\cite{Sta03b}.
247: \section{Scalar fields on static spherically symmetric\\ backgrounds}\label{sec:scal}
248: \subsection{The space-time}
249: Consider a four-dimensional connected spherically symmetric static space-time
250: $(\mathcal{M},g)$. More precisely on $\M$ we assume a time-like action of $\RR$
251: and a space-like action of $SO(3,\RR)$ commuting with it. These
252: actions should be without fixed points, except that at most one $\RR$-orbit
253: is allowed to be $SO(3,\RR)$-fixed, in which case it will be
254: called the time axis. We restrict our attention to the cases where the $SO(3,\RR)$-orbits
255: off the time axis are spheres.
256:
257: \subsection{The coordinate system}
258: Let $T=T(p)$ denote the Killing vector
259: generating the $\RR$-action at $p\in\M$ and let $A=A(p)$ denote the area of the
260: $SO(3,\RR)$-orbit passing through the point $p\in\M$. We define two $\RR\times SO(3,\RR)$-invariant
261: functions on $\M$ as follows:
262: \[
263: \alpha := \sqrt{-g(T,T)},\qquad
264: \rho := \sqrt{A/4\pi}.
265: \]
266: The quotient space $(\mathcal{Q},\overline{g}):=(\mathcal{M},g)/SO(3,\RR)$
267: is a two-dimensional Lorentzian manifold.\footnote{Possibly a
268: manifold with boundary if there is a time axis.}
269: We can assign coordinates $t,r$ on $\mathcal{Q}$ in such a way that
270: \begin{displaymath}
271: \overline{g}_{ab}\,dy^a\,dy^b = \alpha^2(r)\left(-dt^2+dr^2\right).
272: \end{displaymath}
273: The Killing field is clearly just $\partial_t$
274: in these coordinates. We pull back
275: $t$ and $r$ to $\mathcal{M}$. It is easy to see that
276: we can assign angular coordinates $\Omega = (\theta,\phi)$ on $\mathcal{M}$ in
277: such a way that the metric becomes
278: \begin{equation}\label{metricform}
279: g_{\mu\nu}\,dx^\mu\,dx^\nu = \alpha^2(r) \left(-dt^2+dr^2\right)
280: + \rho(r)^2 (d\phi^2+\sin^2\phi d\theta^2),
281: \end{equation}
282: and the action of $SO(3,\RR)$ is the usual action on the unit
283: sphere. These coordinates are unique up to translations in the
284: $t,r$ coordinates and rotations in the $\theta,\phi$ coordinates.
285: If there is a time axis we can then arrange that $r=0$ there.
286:
287: We note that in these coordinates the volume form of the spacetime is
288: \[
289: d\mu_g = \alpha^2 \rho^2 d\Omega dr dt,
290: \]
291: while the induced volume form on the Cauchy hypersurfaces $t = \mbox{const.}$ is
292: \[
293: d\sigma = \alpha \rho^2 d\Omega dr.
294: \]
295: Here $$d\Omega = \sin\phi d\theta d\phi$$ denotes the volume form on the standard 2-sphere. We also note that there are at least two other volume forms on these hypersurfaces that are natural to
296: consider, namely
297: \[
298: d\sigma' = \alpha d\sigma = \alpha^2 \rho^2 d\Omega dr\qquad\mbox{ and }
299: \qquad d\sigma'' = \alpha^{-1} d\sigma = \rho^2 d\Omega dr.
300: \]
301: The significance of $d\sigma'$ is that $L^p$-norms based on it behave as one would
302: expect with respect to the $t$-foliation: For $f:\M \to \RR$
303: and all $p\geq 1$,
304: \[
305: \| f \|_{L^p(d\mu_g)} = \|\ \| f \|_{L^p(d\sigma')} \|_{L^p(dt)}.
306: \]
307: The significance of $d\sigma''$ will become clear in the next section.
308:
309: \subsection{The wave equation}
310: Consider the action functional $\A = \half \int_\M L d\mu_g$ corresponding to a real massless scalar field $\psi:\M\to\RR$. The Lagrangian density is
311: $$L = g^{\mu\nu}\nab_\mu \psi \nab_\nu \psi.$$
312: In the above-given coordinates
313: \[
314: L = -\alpha^{-2} \psi_t^2 + \alpha^{-2} \psi_r^2 + \rho^{-2}|\sphgrad
315: \psi|^2,
316: \]
317: where $\sphgrad$ denotes the unit-spherical gradient, and
318: \[
319: |\sphgrad \psi|^2 = |\partial_\phi \psi|^2 + \frac{1}{\sin^2\phi} |\partial_\theta \phi|^2.
320: \]
321: A scalar field $\psi$ that is a stationary point of the action $\A$, subject to
322: a given set of initial values $(\psi_0,\psi_1)$ on the Cauchy hypersurface $t=0$, satisfies the
323: following Cauchy problem
324: \begin{equation}\label{eq:psi}
325: \partial_t^2 \psi + A \psi = 0,\qquad \psi(0) = \psi_0,\qquad
326: \partial_t\psi(0) = \psi_1,
327: \end{equation}
328: where
329: \begin{equation}\label{def:A}
330: A := - \frac{1}{\rho^2} \partial_r (\rho^2 \partial_r) -
331: \frac{\alpha^2}{\rho^2} \sphlap,
332: \end{equation}
333: with $\sphlap$ denoting the Laplace-Beltrami operator on the unit 2-sphere. Note that the operator $A$ is symmetric
334: and positive definite with respect to the inner product given by the volume form $d\sigma''$,
335: i.e. for $\phi,\psi \in C^\infty_c(\Sigma_t)$,
336: \[
337: \int_{\Sigma_t} \phi A \psi\ d\sigma'' = \int_{\Sigma_t}\psi A\phi\ d\sigma'',
338: \]
339: and
340: \[
341: \int_{\Sigma_t} \phi A \phi\ d\sigma'' =
342: \int_{\Sigma_t} \phi_r^2 + \frac{\alpha^2}{\rho^2} |\sphgrad \phi|^2\
343: d\sigma'' \geq 0,
344: \]
345: with equality only if $\phi\equiv 0$.
346:
347: \section{Self-Adjoint Extensions, Energy, and Sobolev Norms}\label{sec:energy}
348: As explained in Section~\ref{sec:intro}, the evolution
349: problem (\ref{eq:psi}) is only meaningful if the operator $A$ there is replaced
350: by (one of) its self-adjoint extension(s) $A_E$. In the event that the self-adjoint
351: extension is not unique, the
352: one we choose to pick is the Friedrichs extension $A_F$, obtained by extending the
353: quadratic
354: form naturally associated with the operator $A$, namely
355: \begin{equation}\label{def:Q}
356: Q_A(\phi) := \int_{\Sigma_t} \phi_r^2 + \frac{\alpha^2}{\rho^2}
357: |\sphgrad \phi|^2\ d\sigma''.
358: \end{equation}
359: Thus the Cauchy problem we are actually studying is the following
360: \begin{equation}\label{eq:psiFri}
361: \partial_t^2 \psi + A_F \psi = 0,\qquad \psi_{|_\Sigma} = f,\qquad \partial_t \psi_{|_\Sigma} = g.
362: \end{equation}
363:
364: Let $T_{\mu\nu}$ be the energy tensor of the field, defined as
365: \[
366: T_{\mu\nu} = \nab_\mu \psi \nab_\nu \psi - \half g_{\mu\nu}L.
367: \]
368: In particular, in the above given coordinates
369: \[
370: T_{00} = \half \psi_t^2 + \half \psi_r^2 + \frac{\alpha^2}{2\rho^2}
371: |\sphgrad \psi|^2.
372: \]
373: Let $X$ be a time-like Killing vector field for $\M$. Then the one-form
374: $P$ defined by $P(Y) := T(X,Y)$ is divergence-free, i.e. ${}*d*P = 0$ and thus $*P$ is a
375: conserved current.
376: In particular, let
377: $X= \partial_t$, then $P_0 = T_{00}$ and $P_i = T_{0i}$, so that the only nonzero
378: component of $*P$ is
379: $(*P)_{123} = \rho^2 T_{00}$, and the conserved quantity is the {\em energy}
380: \[
381: \mathbf{E}[\psi]:=\int_{\Sigma_{t}} \rho^2 T_{00}\ dr d\Omega =
382: \int_{\Sigma_{t}} T_{00}\ d\sigma''.
383: \]
384: Thus using the $d\sigma''$ volume form, $T_{00}$ is identified with the
385: {\em energy density} of the field.
386:
387: It therefore makes sense to also use $d\sigma''$ to define Sobolev spaces on the $\Sigma_t$'s:
388: Let $\Hi^1(\Sigma_t)$ denote the
389: completion of smooth compactly supported functions on $\Sigma_t$ with respect to the norm
390: \[
391: \| f\|_{\Hi^1}^2 := \int_{\Sigma_t} |f_r|^2 + \frac{\alpha^2}{\rho^2}
392: |\sphgrad f|^2 \ d\sigma''.
393: \]
394: We define $\Hi^0(\Sigma_t)$ in the same way, i.e. completion with respect to the norm
395: \[
396: \| f \|_{\Hi^0}^2 := \int_{\Sigma_t} |f|^2\ d\sigma''.
397: \]
398: The Sobolev spaces $\Hi^s(\Sigma_t)$ for $0<s<1$ are then defined via interpolation between
399: the above
400: two spaces, and one uses duality to define them also for $-1\le s<0$. We note that by these
401: definitions,
402: \[
403: Q_A(\phi) = \| \phi\|_{\Hi^1(\Sigma_t)}^2,
404: \]
405: and moreover
406: \[
407: \mathbf{E}[\psi] = \half \left(\| \psi \|_{\Hi^1(\Sigma_t)}^2 +
408: \| \psi_t \|_{\Hi^0(\Sigma_t)}^2\right).
409: \]
410: More generally, for $0 \le s \le 1$ we can define the $s$-energies
411: \[
412: \mathbf{E}_s[\psi] := \half\left( \|\psi\|_{\Hi^s(\Sigma_t)}^2 +
413: \| \psi_t \|_{\Hi^{s-1}(\Sigma_t)}^2 \right),
414: \]
415: and it is not hard to see that they are all conserved under the flow (\ref{eq:psiFri}):
416: \[
417: \frac{d}{dt} \mathbf{E}_s[\psi] = 0.
418: \]
419:
420: \section{Transferring to Minkowski Space}
421: It is clear that if we can view (\ref{eq:psi}) as an evolution in Minkowski space, we may then be able
422: to apply known theorems in order to get the estimates we want. This is of course only possible if the
423: manifold $\M$ has at least the same topology as $\RR^4$ (or $\RR^4$ minus a line,
424: in order to allow a singular time axis). We will be making this assumption from now on, and denote
425: the coordinates on $\RR^4$ by the same letters as those on $\M$, namely
426: $(t,r,\Omega)\in \RR\times \RR^+\times \BS^2$, and a $t$-slice in $\RR^4$ is still denoted by $\Sigma_t$.
427: Let us define $u:\RR^4\to \RR$ by
428: \[
429: u(t,r,\Omega):= \frac{\rho(r)}{r} \psi(t,r,\Omega).
430: \]
431: Note that
432: \begin{equation}\label{equiL2}
433: \| u \|_{L^2(\Sigma_t)}^2 = \int_{\Sigma_t} |u|^2 r^2 dr d\Omega =
434: \int_{\Sigma_t} |\psi|^2 d\sigma'' = \|\psi\|_{\Hi^0(\Sigma_t)}^2.
435: \end{equation}
436:
437: We can easily check that for $\psi\in C^\infty_c(\Sigma_t)$ we can perform an integration by parts
438: to obtain
439: \begin{eqnarray}
440: Q_A(\psi) &=& \int_{\Sigma_t} \left[ u_r^2 + \frac{\alpha^2}{\rho^2} |\sphgrad u|^2 +
441: \frac{\rho^2}{r^2}[\partial_r(\frac{r}{\rho})]^2 u^2 +
442: \frac{\rho}{r}\partial_r(\frac{r}{\rho})\partial_r (u^2) \right] r^2 dr\,d\Omega\nonumber\\
443: & = & \int_{\Sigma_t} \left[ u_r^2 + \frac{\alpha^2}{\rho^2} |\sphgrad u|^2 +
444: V(r) u^2 \right] r^2 dr d\Omega =: Q_B(u),\label{equiQ}
445: \end{eqnarray}
446: where the ``potential" $V$ is defined by
447: \begin{equation}\label{def:V}
448: V(r) = \rho''(r)/\rho(r).
449: \end{equation}
450: Let $B$ denote the operator whose associated quadratic form is $Q_B$ defined above, i.e.
451: \[
452: B := -\frac{1}{r^2}\partial_r(r^2\partial_r) - \frac{\alpha^2}{\rho^2}
453: \sphlap + V.
454: \]
455: We then have that
456: \[
457: \partial_t^2 + A = \frac{r}{\rho} (\partial_t^2 + B) \frac{\rho}{r}.
458: \]
459: Note that the first term in the definition of $B$ coincides with the radial flat Laplacian.
460: In particular, if $\psi$ satisfying (\ref{eq:psi}) is spherically symmetric, i.e. $\psi = \psi(t,r)$, then
461: $u$ can be viewed as a {\em radial} solution of the flat wave equation with a potential
462: \begin{equation}\label{eq:u}
463: u_{tt} - u_{rr} - \frac{2}{r} u_r + V(r) u = 0,\qquad u(0,r) =
464: f(r),\qquad u_t(0,r) = g(r),
465: \end{equation}
466: with $f = \rho \psi_0/r $ and $g = \rho \psi_1/r$.
467:
468: Since by (\ref{equiQ}) the quadratic forms corresponding to $A$ and $B$ are equivalent,
469: we can identify the Friedrichs extensions $A_F$ and $B_F$
470: of $A$ and $B$, as well
471: as the Sobolev spaces based on them, i.e. let $\Hc^s$ denote
472: the $B$-based Sobolev space, defined to be the completion of smooth compactly supported functions
473: on $\RR^3\setminus\{0\}$ with respect to the norm
474: \[
475: \|f\|_{\Hc^s} := \| (B_F)^{s/2} f \|_{L^2}.
476: \]
477: We then have that for $|s|\leq 1$
478: \[
479: \| u \|_{\Hc^s} = \| \psi \|_{\Hi^s}.
480: \]
481: This is because the $L^2$ spaces agree by (\ref{equiL2}) while the $H^1$ spaces agree because of (\ref{equiQ}).
482: On the other hand, if we let
483: \[
484: P := -\Delta+V = - \frac{1}{r^2}\partial_r(r^2\partial_r) - \frac{1}{r^2} \sphlap +V,
485: \]
486: then it is clear that
487: on the subspace of radial functions, $B_F$ coincides with $P_F$ (the Friedrichs extension of $P$), and so do the
488: corresponding Sobolev spaces, which we will denote by $\Hc_\rad$. In particular if we define the $s$-energy of $u$ to be
489: \[
490: \E_s[u] := \| u(t) \|_{\Hc^s_\rad} + \|u_t(t)\|_{\Hc^{s-1}_\rad},
491: \]
492: then
493: \begin{equation}\label{Eequiv}
494: \E_s[u] = \mathbf{E}_s[\psi]
495: \end{equation}
496: and it is conserved by the flow (\ref{eq:u}).
497:
498: In \cite{BPST2} it was shown that solutions to
499: \begin{equation}\label{eq:P}
500: (\partial_t^2 + P_F)u = 0
501: \end{equation}
502: satisfy certain
503: weighted-$L^2$ (Morawetz) and $L^p$ (Strichartz)
504: spacetime estimates given below, provided the potential $V$ meets certain criteria.
505: In the case of a {\em radial} potential $V =V(r)$ on $\RR^3$ these criteria reduce to the following three conditions
506: on $V$:
507: \begin{eqnarray}\label{cond1}
508: \sup_{r>0} r^2 V(r) & < & \infty,\\
509: \inf_{r>0} r^2 V(r)& > & -1/4, \label{cond2}\\
510: \sup_{r>0} r^2 \frac{d}{dr}(r V(r)) &<& 1/4. \label{cond3}
511: \end{eqnarray}
512: The version of the main result in \cite{BPST2} for radial potentials (and general data) is as follows:
513: \begin{theorem}\label{thm:bpst2}
514: Let $V\in C^1((0,\infty),\RR)$ satisfy (\ref{cond1},\ref{cond2},\ref{cond3}), and let $P := -\Delta + V(|x|)$
515: where $\Delta$ is the Laplace operator with domain $C^\infty_c(\RR^3\setminus\{0\})$, and let $P_F$ denote its Friedrichs extension.
516: Then
517: there exists a constant $C$, depending only on the quantities on the
518: left in (\ref{cond1},\ref{cond2},\ref{cond3}), such that any
519: solution $u$ of (\ref{eq:P}) satisfies
520: \begin{equation}\label{est:morstr}
521: \| r^{-1} u \|_{L^2(\RR^4)} + \| u\|_{L^4(\RR^{4})} \leq C \E_{1/2}[u].
522: \end{equation}
523: \end{theorem}
524:
525: Now, the space-time $L^p$ norms of $\psi$ and $u$ are related in the following way:
526: \[
527: \|\psi\|_{L^p(\M)}^p = \int_\M |\psi|^p d\mu_g =
528: \int_{-\infty}^\infty\int_{\BS^2}\int_0^\infty \alpha^2
529: \left(\frac{r}{\rho}\right)^{p-2}\ |u|^p \ r^2 dr\, d\Omega\, dt.
530: \]
531: In particular
532: \begin{equation}\label{L4equiv}
533: \|\psi\|_{L^4(\M)}^4 = \int_{\RR^4} \left(\frac{\alpha r}{\rho}\right)^2
534: |u|^4 d^3x dt.
535: \end{equation}
536: Similarly
537: \begin{equation}\label{L2equiv}
538: \| \frac{\psi}{\rho}\|_{L^2(\M)}^2 = \int_{\RR^4} \left(\frac{\alpha r}{\rho}\right)^2
539: \frac{|u|^2}{r^2}
540: d^3x dt.
541: \end{equation}
542:
543: We thus have proved the following theorem regarding estimates for the scalar field
544: $\psi$ on $\M$:
545:
546: \begin{theorem}\label{thm:main}
547: Let $\M$ be a Lorentzian manifold that is homeomorphic to~$\RR^4$, admitting a timelike $\RR$
548: action and a spacelike $SO(3,\RR)$ action commuting with it, in such a way that exactly one
549: $\RR$-orbit, called $\Gamma$, is $SO(3,\RR)$-fixed. Let
550: $(t,r,\Omega)\in \RR\times\RR^+\times \BS^2$ be the coordinate system on $\M$ as in
551: Section~\ref{sec:scal}, with $\Gamma = \{r=0\}$. Let $g$ be a Lorentzian metric
552: on $\M$ that is of
553: the form (\ref{metricform}), is $C^3$ outside $\Gamma$,
554: and is such that the functions $\rho$ and $\alpha$ satisfy
555: the following conditions
556: \begin{itemize}
557: \item[\rm(i)] $\sup_{r>0} (r^2V) < \infty$
558: \item[\rm(ii)] $\inf_{r>0} (r^2 V) > -1/4$
559: \item[\rm(iii)] $\sup_{r>0} (r^2\frac{d}{dr}(rV)) < 1/4$
560: \item[\rm(iv)] $\inf_{r>0} (\ds\frac{\rho}{\alpha r}) > 0$
561: \end{itemize}
562: where
563: \[
564: V(r) := \frac{1}{\rho}\frac{d^2\rho}{dr^2}.
565: \]
566: Then there exists a constant $C>0$, depending only on the quantities on the
567: left in the conditions above,
568: such that any spherically symmetric solution of \[ \partial_t^2 \psi + A_F \psi = 0 \] satisfies
569: \begin{equation}\label{est:main}
570: \| \rho^{-1} \psi\|_{L^2(\M)} + \| \psi\|_{L^4(\M)} \leq C \mathbf{E}_{1/2}[\psi].
571: \end{equation}
572: \end{theorem}
573:
574: {\em Proof:} We simply observe that by (\ref{L4equiv},\ref{L2equiv}),
575: \[
576: \| \rho^{-1}\psi\|_{L^2} + \| \psi \|_{L^4} \leq \frac{1}{d} \|r^{-1} u \|_{L^2} + \frac{1}{\sqrt{d}} \| u \|_{L^4}
577: \]
578: where $d$ denotes the quantity on the left in condition (iv). The result then
579: follows from (\ref{est:morstr}) and (\ref{Eequiv}).
580: \qed
581:
582: \section{Super-extremal Reissner-Nordstr\"om}
583:
584: The Reissner-Nordstr\"om manifold $(\M,g)$ is the static, spherically
585: symmetric solution of Einstein-Maxwell equations. It is characterized
586: by two parameters: mass $m$ and charge $e$.
587: It can be shown that for a metric of the form (\ref{metricform}) to satisfy the Einstein-Maxwell system, one must have
588: \begin{equation}\label{ode}
589: \frac{d\rho}{dr} = \alpha^2
590: \end{equation}
591: and
592: \begin{equation}\label{def:alpha}
593: \alpha = \sqrt{1 - \frac{2m}{\rho} + \frac{e^2}{\rho^2}}.
594: \end{equation}
595: It thus follows that when $|e|<m$ the function $\rho$ is bounded away from zero and there is no time axis. This is called the sub-extremal (black hole) case.
596: When $|e|>m$ the manifold $\M$ has the same topology as $\RR^4$ minus a line,
597: the time axis is at $r=0$ and the metric $g$ is highly singular there. This is the super-extremal (naked) case of Reissner-Nordstr\"om.
598:
599: The ODE (\ref{ode}) can be solved to express the ``tortoise" coordinate $r$ as a function of $\rho$. In the super-extremal case,
600: \begin{equation}\label{longugly}
601: r(\rho) = r_0 + \rho + m\log(\frac{\rho^2-2m\rho+e^2}{e^2-m^2}) +
602: \frac{2m^2-e^2}{\sqrt{e^2-m^2}}\tan^{-1}\frac{\rho-m}{\sqrt{e^2-m^2}}.
603: \end{equation}
604: We choose the constant $r_0$ such that $r(0) =0$. Since $r$ is an increasing function of $\rho$ this
605: implicitly defines $\rho$ as a function of $r$.
606: It is also easy to compute from the ODE (\ref{ode}) that
607: \[
608: \lim_{r\to 0} \frac{\rho}{r^{1/3}} = (3e^2)^{1/3},\qquad \lim_{r\to \infty} \frac{\rho}{r} = 1.
609: \]
610: and as a result
611: \[
612: \lim_{r\to 0} r^{1/3} \alpha = (e/3)^{1/3},\qquad \lim_{r\to \infty} \alpha = 1,
613: \]
614: so that condition (iv) of Theorem~\ref{thm:main} is clearly satisfied since
615: \[ \frac{\rho}{\alpha r} \to \infty\mbox{ as }r\to 0,\qquad \frac{\rho}{\alpha r} \to
616: 1 \mbox{ as } r \to \infty.
617: \]
618:
619: To prove Theorem~\ref{thm:intro}, it thus remains to check that the function
620: \[
621: V = \frac{1}{\rho}\frac{d^2\rho}{dr^2} =
622: \frac{2m}{\rho^3} - \frac{2e^2 + 4m^2}{\rho^4}
623: + \frac{6me^2}{\rho^5} - \frac{2e^4}{\rho^6},
624: \]
625: satisfies conditions (i-iii) of Theorem~\ref{thm:main}. In the case of condition (iii) this does not appear to be
626: an easy task, because of the
627: transcendental relation (\ref{longugly}) between $r$ and $\rho$, and the dependence on the two additional variables
628: $e$ and $m$. To overcome these difficulties, first we exploit the inherent scaling in the problem to
629: eliminate $e$ (effectively setting it equal to one)
630: and then by forgetting the relationship between $r$ and $\rho$ and treating them as independent quantities,
631: we transform the problem into one about
632: polynomials in three variables, which we then solve by utilizing a few basic tools from the theory of
633: real algebraic curves.
634:
635: \begin{remark}\rm The lower bound on the charge-to-mass ratio in Theorem~\ref{thm:intro} is not sharp. The actual
636: ratio for which condition (iii) of Theorem~\ref{thm:main} starts to be violated is less than 2 (It is
637: not hard to see that this condition is violated for the ratios close to 1). We have not
638: attempted to find the precise cut-off point, since it is not known whether condition (iii) is necessary
639: for the estimate (\ref{est:main}) to hold. We also note that
640: the charge-to-mass ratios that occur in nature, such as in
641: elementary particles, are in fact very large\footnote{The charge $e$ and mass $m$ appearing in
642: (\ref{def:alpha}) are
643: in geometric rationalized units, with $G = c = \hbar = 1$. The corresponding quantities in unrationalized
644: electrostatic units $e^*$ and $m^*$ are related to $e$ and $m$ in the following way: $m = \frac{G}{c^2} m^*$ and
645: $e = \frac{\sqrt{G}}{c^2} e^*$. It follows that $e/m \approx 10^{21}$ for an electron and $\approx 2\times 10^{18}$ for a proton. }.
646: \end{remark}
647:
648: We perform the scaling first: Let
649: \[
650: x := \frac{r}{|e|},\qquad y := \frac{\rho}{|e|},\qquad z := \frac{m}{|e|},\qquad v := e^2 V
651: \]
652: We then have
653: \[
654: \alpha^2 = 1 - \frac{2z}{y} + \frac{1}{y^2}.
655: \]
656: Let $y = \phi_z(x)$ denote the solution to the following initial value problem
657: \[
658: \frac{dy}{dx} = \alpha^2(y,z),\quad y(0) = 0.
659: \]
660: The inverse function of $\phi_z$ is in fact explicit:
661: \[
662: \phi_z^{-1}(y) = y + \log(y^2 - 2 yz +1) +
663: \frac{2z^2-1}{\sqrt{1-z^2}}\tan^{-1}\frac{y\sqrt{1-z^2}}{1-yz}.
664: \]
665: Note that since we are considering the super-extremal case, the parameter~$z$ ranges from 0 to 1.
666:
667: From the ODE it is clear that
668: \begin{equation}\label{phiasymp}
669: \lim_{x\to 0} \frac{(\phi_z(x))^3}{x} = 3,\qquad \lim_{x\to \infty}\frac{\phi_z(x)}{x} = 1.
670: \end{equation}
671: We have
672: \[
673: v(y,z) = \frac{1}{y} \frac{d^2 y}{dx^2} = \frac{2z}{y^3}-\frac{2+4z^2}{y^4} + \frac{6z}{y^5}-\frac{2}{y^6} =
674: \frac{2}{y^6}(yz-1)(y^2-2yz+1).
675: \]
676: From (\ref{phiasymp}) it follows that
677: \begin{eqnarray}
678: \lim_{x\to 0}x^2 v(\phi_z(x),z) & = & -\frac{2}{9},\label{vasmyp0}\\
679: \lim_{x\to\infty} x^2 v(\phi_z(x),z) & = & 0,\label{vasympinf}
680: \end{eqnarray}
681: and thus conditions (i) and (ii) of Theorem~\ref{thm:main} are satisfied since
682: $-\frac{2}{9}>-\frac{1}{4}$ and $v$ is clearly
683: negative and increasing for $y<1/z$, positive for $y>1/z$, and smooth away from $x=0$.
684:
685: \begin{remark}\label{rem:hair}\rm
686: Note however that since $-\frac{2}{9}<\frac{3}{4}$, the operator
687: $$Q:= -\p_x^2 -\frac{2}{x}\p_x + v(\phi_z(x),z)$$ with domain $C^\infty_c(\RR^+)$ will have non-unique self-adjoint extensions, i.e. is
688: limit-circle at zero (see \cite[Appendix to X.1]{ReeSimII}). In this sense, with respect to {\em spherically symmetric} scalar waves, the naked singularity at $x=0$ of the super-extremal Reissner-Nordstr\"om
689: solution has a $U(1)$ hair. Note also that the angular momentum contribution to the full Laplace-Beltrami
690: operator of this manifold is
691: \[
692: \frac{\alpha^2(\phi_z(x),z)}{(\phi_z(x))^2}\ell(\ell+1)
693: \]
694: where $\ell$ is the spherical harmonic degree. Since this behaves like $x^{-4/3}$ near
695: the origin, its
696: addition to $Q$ does not change anything as far as self-adjoint extensions are concerned.
697: It thus follows that with respect to scalar waves in general, this naked singularity is
698: ``infinitely hairy," i.e. the corresponding unitary group is certainly infinite-dimensional\footnote{This provides us with another reason for choosing the Friedrichs extension instead.}.
699: \end{remark}
700: Consider now the expression in condition (iii) of Theorem~\ref{thm:main}:
701: \begin{eqnarray*}
702: x^2\frac{d}{dx}(xv(\phi_z(x),z)) &=&
703: x^2v(\phi_z(x),z)+ x^3 \frac{\partial v}{\partial y}(\phi_z(x),z)
704: \alpha^2(\phi_z(x),z)\\
705: & = & w(x,\phi_z(x),z),
706: \end{eqnarray*}
707: where
708: \[
709: \begin{array}{rcl}
710: w(x,y,z) &:=& \left(-\frac{6z}{y^4} + \frac{8 + 28z^2}{y^5}
711: - \frac{52z + 32z^3}{y^6} + \frac{20 + 76z^2}{y^7}
712: - \frac{54z}{y^8} + \frac{12}{y^9}\right)x^3
713: \cr && {}
714: + \left(\frac{2z}{y^3} - \frac{2 + 4 z^2}{y^4}
715: + \frac{6z}{y^5} - \frac{2}{y^6}\right)x^2.
716: \end{array}
717: \]
718: Let us define the function $j_z:\RR^+\to\RR^+$ by
719: \begin{equation}\label{def:jz}
720: j_z(y) = w(\phi_z^{-1}(y),y,z).
721: \end{equation}
722: A long, routine computation of Taylor series shows that near $y=0$,
723: \begin{equation}\label{taylor}
724: j_z(y) = \frac{2}{9} - \frac{4-z^2}{270} y^2 + O(y^3),
725: \end{equation}
726: \begin{equation}\label{taylor2}
727: j'_z(y) = - \frac{4-z^2}{135} y + O(y^2),
728: \end{equation}
729: and
730: \begin{equation}\label{taylor3}
731: j''_z(y) = - \frac{4-z^2}{135} + O(y).
732: \end{equation}
733: while, for large $y$,
734: \begin{equation}\label{taylor4}
735: j_z(y) = -4zy^{-3} - 28zy^{-4}\log y + O(y^{-4})
736: \end{equation}
737: and
738: \begin{equation}\label{taylor5}
739: j'_{z}(y) = 12zy^{-4} + 112zy^{-5}\log y + O(y^{-5}).
740: \end{equation}
741: The implied constants in $O$-terms are locally uniform in~$z$.
742: Thus $j_z$ has a local maximum at $y=0$, for $0\leq z<1$.
743: We wish to prove that, for $z \leq 1/2$, this is in fact the global maximum of $j_z$. Since it is easy to see
744: that $j_z\to 0$ as
745: $y\to\infty$, it suffices to show that the only local maximum is the
746: one at $y=0$. To this end we first show that
747: \begin{prop}\label{prop:nospi}
748: $j_z$ has no stationary
749: points of inflection as long as $z\leq 1/2$.
750: \end{prop}
751:
752: {\em Proof:} We compute
753: \[
754: j'_z(y) = \frac{\partial w}{\partial y}(\phi_z^{-1}(y),y,z) + \frac{\partial w}{\partial x}(\phi_z^{-1}(y),y,z)\frac{1}{\alpha^2(y,z)} = q_1(\phi_z^{-1}(y),y,z),
755: \]
756: where
757: \[
758: \begin{array}{rcl}
759: q_1(x,y,z) & := & \left( {\frac {24 z}{{y}^{5}}}-{\frac {40+140\,{z}^{2}}{{y}^{6}}}+{
760: \frac {312\,z+192\,{z}^{3}}{{y}^{7}}}\right.\\
761: &&\quad-\left. {\frac {140+532\,{z}^{2}}{{y}^{
762: 8}}}+{\frac {432 z}{{y}^{9}}}-\frac{108}{y^{10}} \right) {x}^{3}\\
763: &&{} + \left(
764: -{\frac {24z}{{y}^{4}}}+{\frac {64\,{z}^{2}+32}{{y}^{5}}}-\,{
765: \frac {120 z}{{y}^{6}}}+\frac{48}{y^{7}} \right) {x}^{2}\\
766: &&{} + \left( {\frac {4z}
767: {{y}^{3}}}-\frac{4}{{y}^{4}} \right) x,
768: \end{array}
769: \]
770: and similarly we compute
771: \[
772: j_z''(y) = \frac{1}{\alpha^2(y,z)} q_2(\phi_z^{-1}(y),y,z),
773: \]
774: where
775: \[
776: \begin{array}{rcl}
777: q_2(x,y,z) & := &
778: \left( -{\frac {120 z}{{y}^{4}}}+{\frac {240+1080\,{z}^{2}}{{y}^{5}}}
779: +{\frac {-3024\,{z}^{3}-2784\,z}{{y}^{6}}}\right.\\
780: &&\quad\left. +{\frac {1360+9464\,{z}^{2}
781: +2688\,{z}^{4}}{{y}^{7}}}+{\frac {-8312\,z-9856\,{z}^{3}}{{y}^{8}}}\right.\\
782: &&\quad+\left. {
783: \frac {2200+12032\,{z}^{2}}{{y}^{9}}}-\,{\frac {6048z}{{y}^{10}}}+\frac{1080}
784: {y^{11}} \right) {x}^{3}\\
785: &&{} + \left( {\frac {168 z}{{y}^{3}}}+{\frac {
786: -932\,{z}^{2}-280}{{y}^{4}}}+{\frac {2072\,z+1216\,{z}^{3}}{{y}^{5}}}\right.\\
787: &&\quad\left. +
788: {\frac {-3356\,{z}^{2}-916}{{y}^{6}}}+\,{\frac {2688z}{{y}^{7}}}-\frac{660}
789: {y^{8}} \right) {x}^{2}\\
790: &&{} + \left( -{\frac {60 z}{{y}^{2}}}+{\frac {80+
791: 152\,{z}^{2}}{{y}^{3}}}-\,{\frac {284z}{{y}^{4}}}+\frac{112}{y^{5}}
792: \right) x\\
793: &&{}+{\frac {4z}{y}}-\frac{4}{y^{2}}.
794: \end{array}
795: \]
796: To prove the statement of the proposition, it is enough to show that the zero-sets of
797: $q_1(x,y,z)$ and $q_2(x,y,z)$ are disjoint. The strategy is to forget about the
798: transcendental relationship between $x$ and $y$ and treat them as independent variables,
799: in order to be able to use results from the theory of real algebraic curves.
800:
801: At a stationary point of inflection both $j'_z$ and $j''_z$ would vanish
802: and hence so would the resultant\footnote{By definition, the resultant
803: of two polynomials
804: \[
805: q_1 = a_n \Pi_{i=1}^n (x-\alpha_i),\qquad q_2 = b_m \Pi_{i=1}^m (x-\beta_i)
806: \]
807: is
808: \[
809: R(q_1,q_2;x) = a_n^m b_m^n \Pi_{i=1}^n\Pi_{j=1}^m (\alpha_i - \beta_j)
810: \]
811: It can be computed from the Euclidean algorithm, or as the determinant of Sylvester's matrix
812: or Bezout's matrix.} of $q_1,q_2$,
813: considering both as cubic polynomials
814: in~$x$.
815: Up to an integer multiple, this resultant is computed\footnote{The necessary
816: computations here and elsewhere in the paper are performed by the computer algebra package
817: PARI using exact integer arithmetic.} to be
818: \[
819: R(q_1,q_2;x) =
820: y^{-34} (yz-1)^2
821: p(y,z),
822: \]
823: where
824: \begin{equation}\label{def:p}
825: \begin{array}{rcl}
826: p(y,z) &=& \left(-1536\,{y}^{9}+4608\,{y}^{7} \right) {z}^{9}\\
827: &&{}+ \left( 3452\,{y}^{10}+2040\,{y}^{8}-20100\,{y}^{6} \right) {z}^{8}\\
828: &&{}+ \left( -3504\,{y}^{
829: 11}-16456\,{y}^{9}-8608\,{y}^{7}+36120\,{y}^{5} \right) {z}^{7}\\
830: &&{}+
831: \left( 1947\,{y}^{12}+20360\,{y}^{10}+62966\,{y}^{8}+48272\,{y}^{6}-
832: 33769\,{y}^{4} \right) {z}^{6}\\
833: &&{}+ \left( -576\,{y}^{13}-11988\,{y}^{11}-
834: 71800\,{y}^{9}-153832\,{y}^{7}\right.\\
835: &&\qquad -\left. 104440\,{y}^{5}+17900\,{y}^{3} \right)
836: {z}^{5}\\
837: &&{}+ \left( 72\,{y}^{14}+3552\,{y}^{12}+38762\,{y}^{10}+143492\,{
838: y}^{8}\right.\\
839: &&\qquad+\left. 208760\,{y}^{6}+109100\,{y}^{4}-5530\,{y}^{2} \right) {z}^{4}\\
840: &&{}+
841: \left( -432\,{y}^{13}+10464\,{y}^{11}-66316\,{y}^{9}-154672\,{y}^{7}\right. \\
842: &&\qquad-\left.
843: 151552\,{y}^{5}-62848\,{y}^{3}+972\,y \right) {z}^{3}\\
844: &&{}+ \left( 1152\,{
845: y}^{12}+15384\,{y}^{10}+57803\,{y}^{8}+83120\,{y}^{6}\right. \\
846: &&\qquad+\left. 58958\,{y}^{4}+
847: 20912\,{y}^{2}-81 \right) {z}^{2}\\
848: &&{}+ \left( -1440\,{y}^{11}-10536\,{y}^{9
849: }-21648\,{y}^{7}-20824\,{y}^{5}\right. \\
850: &&\qquad-\left. 11680\,{y}^{3}-3888\,y \right) z\\
851: &&{}+720\,
852: {y}^{10}+2160\,{y}^{8}+2760\,{y}^{6}+1908\,{y}^{4}+944\,{y}^{2}+324.
853: \end{array}
854: \end{equation}
855: Note that
856: \[
857: q_1(x,1/z,z) = -8x^2(z^2-1)z^5\left(2xz^3-2xz+1\right),
858: \]
859: so it is zero only for $x=0$ or
860: $x = \frac{1}{2z-2z^3}$.
861: However,
862: \[
863: q_2(1/(2z-2z^3),1/z,z)={\frac { \left( 5\,{z}^{2}-3 \right) {z}^{2}}{1-{z}^{2}}}.
864: \]
865: $z=0$ would correspond to $x=y=\infty$, which is not of interest, and
866: $\sqrt{3/5}>\half$. It is thus enough to show that $p(y,z)$ has no real
867: zeros for $0\leq z \leq \half$. This is easy to establish for $z=0$,
868: since
869: \begin{equation}\label{p0neg}
870: p(y,0) = 720\,{y}^{10}+2160\,{y}^{8}+2760\,{y}^{6}+1908\,{y}^{4}+944\,{y}^{2}+324 > 0
871: \end{equation}
872: for all $y$.
873: We postpone the rest of the proof of this proposition until the end of the section, and instead show first how this implies
874: the desired result about $V$, i.e. that it satisfies condition (iii) of Theorem~\ref{thm:main}.
875: \begin{lemma} The function $j_0$ has no critical points other than at $y=0$. \end{lemma}
876:
877: {\em Proof:} We have for $z=0$
878: \begin{equation}\label{atan}
879: \phi_0^{-1}(y) = y - \tan^{-1} y
880: \end{equation}
881: and
882: \[
883: q_1(x,y,0) = -4 y^{-10}x[(10y^4+35y^2+27)x^2-4(3y^3+2y^5)x+y^6].
884: \]
885: Consider the algebraic curve $q_1(x,y,0) = 0$. It has a branch $x=0$.
886: The other two branches can be transformed into a hyperbola by changing variables to
887: \[
888: \xi := \frac{x}{y^3},\qquad \eta := \frac{x}{y}.
889: \]
890: Let $\mathfrak{H}$ denote the piece of this hyperbola that lies in the first quadrant of the
891: $(\xi,\eta)$ plane, i.e.
892: $${\mathfrak H} = \{(\xi,\eta)\ |\ \xi\geq 0, \eta\geq 0, h(\xi,\eta) = 0\},$$
893: with
894: \begin{equation}\label{hyper}
895: h := 10\eta^2+35 \xi\eta + 27 \xi^2 - 12\xi-8\eta +1.
896: \end{equation}
897: Likewise, let $\mathfrak{T}$ denote the part of the transcendental curve $x = \phi_0^{-1}(y)$
898: in the same quadrant:
899: $${\mathfrak T}=\{(\xi,\eta)\ |\ \xi\geq 0, \eta\geq 0, \tau(\xi,\eta) = 0\},$$
900: where $\tau$ has the following expansion
901: \[
902: \tau = -1 + \frac{1}{3\xi} - \frac{1}{5\xi^2}\eta + \frac{1}{7\xi^3}\eta^2 -\dots
903: \]
904: Recall that we have already shown by (\ref{p0neg}) that the function $j_0$ cannot have a stationary point of inflection.
905: Therefore on each branch of the hyperbola $\mathfrak H$ the second derivative $j_0''$ must be of one
906: sign. Thus one branch must correspond to minima and the other to maxima of $j_0$.
907:
908: It is not hard to see that $\mathfrak{H}_L$, the lower branch of the hyperbola $\mathfrak{H}$, is well separated from ${\mathfrak T}$: Let
909: $R$ be the rectangle $[0,1/9]\times[0,6/37]$ in the $(\xi,\eta)$ plane. Then
910: \begin{claim} $\mathfrak{H}_L$ is contained in $R$ while $\mathfrak{T}\cap R = \emptyset$. \end{claim}
911: {\em Proof:} The $\xi$-intercept of $\mathfrak{H}_L$ is at $\xi=1/9$. The tangent line to $\mathfrak{H}_L$ at the point $(1/9,0)$ is $\eta = -\frac{54}{37}(\xi-\frac{1}{9})$. Since $\mathfrak{H}_L$ is concave down, it lies below this tangent, and the $\eta$-intercept of the tangent line is at $\eta = 6/37$ which shows the first part of the claim. Consider next the ODE satisfied by the curve
912: $x = y - \tan^{-1} y$, i.e. $\frac{dx}{dy} = \frac{y^2}{y^2+1}$. For $y\leq 1$ we have
913: $\frac{dx}{dy} \geq \frac{y^2}{2}$ which upon integration yields
914: \begin{equation}\label{xibnd}
915: x(y) \geq \frac{y^3}{6},
916: \end{equation}
917: and thus $\xi \geq \frac{1}{6}$,
918: while if $y \geq 1$ then $\frac{dx}{dy} \geq \frac{1}{2}$ and thus, using the bound (\ref{xibnd}) at $y=1$,
919: \[
920: x(y) \geq \frac{1}{2}(y - 1) + x(1) \geq \frac{1}{2}(\frac{1}{3}y + \frac{2}{3}y) -
921: \frac{1}{3} \geq \frac{1}{6}y,
922: \]
923: and thus $\eta \geq \frac{1}{6}$. Thus along $\mathfrak{T}$ we have $\min(\xi,\eta)\geq 1/6$, which proves the second part of the claim.
924:
925: Let $\mathfrak{H}_U$ denote the upper branch of the hyperbola $\mathfrak{H}$. It intersects $\mathfrak T$ at $\xi = \frac{1}{3}, \eta = 0$. This point corresponds
926: to the origin of the $(x,y)$ plane where we know that $j_z$ has a local maximum. We have thus
927: shown in the above Claim that $j_0'$ cannot have any local minima. Moreover, we can compute the tangent lines
928: to the two curves at $(1/3,0)$ and we obtain that
929: \[
930: \left(\frac{d\eta}{d\xi}\right)_{\mathfrak{H}_U} = -\frac{18}{11},\qquad
931: \left(\frac{d\eta}{d\xi}\right)_{\mathfrak T} = -\frac{5}{3}.
932: \]
933: Thus $\mathfrak{H}_U$ is below $\mathfrak T$ near $(\frac{1}{3},0)$. It is easy to see that this is also
934: the case near the $\eta$ axis (which corresponds to the infinity of the $(x,y)$ plane). Thus if
935: $\mathfrak T$ were to intersect $\mathfrak{H}_U$ it would have to do so at least twice. Since a continuous
936: function cannot have two consecutive local maxima without a minimum in between
937: this is not possible and the curves $\mathfrak{H}_U$ and $\mathfrak T$ have no
938: other common point in (the
939: first quadrant of) the $(\xi,\eta)$ plane. This proves the lemma.
940: \qed
941:
942: The following Proposition completes the proof of Theorem~\ref{thm:intro} by showing that the quantity on the left in condition (iii) of Theorem~\ref{thm:main} is equal to $\frac{2}{9}$.
943:
944: \begin{prop}\label{prop:jz} The function $j_z$ for $0\leq z\leq 1/2$ has no local maximum other than at $y=0$. \end{prop}
945: {\em Proof:}
946: Let
947: \[
948: K := \{(y,z)| y > 0,\ j'_z(y) = 0,\ j''_z(y) \leq 0\}.
949: \]
950: Note that $K$ is closed.
951: From (\ref{taylor2}) we see that there is an $\epsilon >0$ such that there
952: $j'_z \neq 0$ in the region $0 < y < \epsilon$, $0 \le z \le \frac 1 2$.
953: Similarly, from (\ref{taylor5}) we see that there is a $Y$ such that
954: $j'_z \neq 0$ in the region $Y < y < \infty$, $0 \le z \le \frac 1 2$.
955: $K$ is therefore compact, since it
956: is a closed subset of $[\epsilon,Y]\times[0,\half]$.
957:
958: If $K$ is nonempty, then there exists a point $(\bar{y},\bar{z})$
959: in $K$ with smallest $z$, i.e.
960: $z\geq \bar{z}$ for all $(y,z)\in K$. Thus
961: $j'_{\bar{z}}(\bar{y}) = 0$ and $j''_{\bar{z}}(\bar{y}) \leq 0$.
962: By the previous lemma, $\bar{z} > 0$. Suppose that
963: $j''_{\bar{z}}(\bar{y}) \ne 0$. Then by the implicit function theorem,
964: we can solve
965: \[
966: q_1(\phi_z^{-1}(y),y,z) = 0
967: \]
968: near $(\bar{y},\bar{z})$, to get the curve $y = \vartheta(z)$.
969: Now for $z<\bar{z}$, $(\vartheta(z),z)\notin K$ so we have
970: $j''_z(\vartheta(z))>0$.
971: Thus by continuity, $j''_{\bar{z}}(\bar{y}) \geq 0$,
972: which leads to a contradiction. Thus we must have
973: $j''_{\bar{z}}(\bar{y}) = 0$.
974: But this is ruled out by Proposition~\ref{prop:nospi}.
975: Therefore $K$ must be empty. This proves the statement of Proposition~\ref{prop:jz}.
976: \qed
977:
978: We conclude with the remainder of the proof of Proposition~\ref{prop:nospi}.
979: First we need some definitions: For
980: \[
981: p(y,z) = \sum_{k,l} c_{k,l} y^k z^l
982: \]
983: a polynomial in two variables with real coefficients, let $C$ be the curve
984: \[
985: C := \{(y,z) \in \RR^2 \ |\ p(y,z) = 0\}
986: \]
987: considered as a subset of $\RR^2$ in the usual topology. The {\em Newton Polygon} of $p$ is defined to be the convex hull of the set
988: \[
989: N := \{(k,l) \in \ZZ^2\ |\ c_{k,l} \ne 0\}.
990: \]
991: If $E$ is an (oriented) edge of the Newton polygon with endpoints $(k'_E,l'_E)$ and $(k''_E,l''_E)$, then the numbers
992: $d_E$, $p_E$ and $q_E$ are defined by
993: \[
994: d_E := \gcd(k''_E-k'_E,l''_E-l'_E),\quad p_E := \frac{k''_E-k'_E}{d_E}, \quad q_E := \frac{l''_E-l'_E}{d_E},
995: \]
996: and the {\em edge polynomial} $e_E \in \RR[t]$ by
997: \[
998: e_E(t) :=\sum_{i=0}^{d_E} c_{k'_E+ip_E,l'_E+iq_E}t^i.
999: \]
1000: An edge is called {\em outer} if it maximizes some linear function $ak+bl$ on the Newton polygon, where at
1001: least one of $a$ or $b$ is positive.
1002:
1003: The following compactness criteria for plane algebraic curves is proved in \cite{Sta03b}:
1004: \begin{theorem}
1005: For the compactness of $C$ it suffices that $p$ is not divisible by $y$ or $z$, and the
1006: edge polynomials corresponding to outer edges have no real zeros.
1007: \end{theorem}
1008:
1009: The Newton polygon of the polynomial (\ref{def:p}) is shown in Figure~\ref{fig:np}.
1010: \begin{figure}[h]
1011: \begin{center}
1012: \includegraphics[width=0.5\textwidth]{nq.eps}
1013: \end{center}
1014: \caption{\label{fig:np} Newton polygon of $p$}
1015: \end{figure}
1016: It is obvious that $p$ does not satisfy the criteria in this theorem, since some of
1017: the outer edge polynomials are of odd degree. To remedy this, we take a simple projective
1018: transformation of the plane and consider its composition with $p$: Let
1019: \begin{equation}\label{def:q}
1020: q(z',y') := {z'}^{18}p(\frac{y'}{z'},\frac{1}{z'}).
1021: \end{equation}
1022: We need to show that $q$ has no zeros in the region $[2,\infty)\times(0,\infty)$.
1023: The Newton polygon of $q$ is the image of the Newton polygon of $p$ under the linear map $(k,l) \mapsto (18-k-l,k)$. It is shown in Figure~\ref{fig:nq}.
1024:
1025: \begin{figure}[h]\begin{center}
1026: \includegraphics[width=0.5\textwidth]{np.eps}
1027: \end{center}\caption{\label{fig:nq} The Newton Polygon of $q$}
1028: \end{figure}
1029:
1030: There are two outer edges, with corresponding edge polynomials
1031: \begin{eqnarray}
1032: e_1(t) & = & 324 t^{10} + 994 t^8 + 1908 t^6 + 2760 t^4 + 2160 t^2 + 720,\\
1033: e_2(t) & = & 72 t^4 - 432 t^3 + 1152 t^2 - 1440 t + 720.
1034: \end{eqnarray}
1035: $e_1$ is clearly never zero, and
1036: the fact that $e_2$ has no real zeros is easily checked by Sturm's criterion.
1037: By the above Theorem,
1038: the zeroes of $q$ are a compact set. If
1039: $z'$ is maximal for this set then $\partial q/\partial y'$ is zero.
1040: The resultant of $q$ and $\partial q/\partial y'$ with respect to $y'$
1041: must then be zero as well.
1042: This resultant is
1043: \[
1044: R(q,\frac{\partial q}{\partial y'};y') =
1045: z'^{114}({z'}-1)^{45}({z'}+1)^{45}(2{z'}-1)(2{z'}+1)f({z'})^3g({z'})
1046: \]
1047: where
1048: \[
1049: \begin{array}{rcl}
1050: f({z'}) &=& 15268608{z'}^{12} - 91375200{z'}^{10} + 235796896{z'}^8
1051: - 336360313{z'}^6\\
1052: && {}
1053: + 278925810{z'}^4 - 126777721{z'}^2 + 24542656
1054: \end{array}
1055: \]
1056:
1057: and
1058: \[
1059: \begin{array}{rcl}
1060: \lefteqn{g({z'})=}&&\\
1061: 2238642500162400000000{z'}^{32}
1062: &-&34444148848863120000000{z'}^{30}
1063: \cr
1064: + 237592851413120362800000{z'}^{28}
1065: &-& 985391370893335206960000{z'}^{26}
1066: \cr
1067: + 2763859141396512532788000{z'}^{24}
1068: &-& 5568163844214174878032000{z'}^{22}
1069: \cr
1070: + 8328390177670642537641736{z'}^{20}
1071: &-& 9407814473561334395291936{z'}^{18}
1072: \cr
1073: + 8074051365793494550609047{z'}^{16}
1074: &-& 5249226125431353947046641{z'}^{14}
1075: \cr
1076: + 2557646890465822492261299{z'}^{12}
1077: &-& 918262815118642547577717{z'}^{10}
1078: \cr
1079: + 238729228492213678314693{z'}^8
1080: &-& 44988866247608231183315{z'}^6
1081: \cr
1082: + 6473535589377087487753{z'}^4
1083: &-& 739775558130688228967{z'}^2
1084: \cr
1085: {}+ 49586421\lefteqn{501845352448.}&&
1086: \end{array}
1087: \]
1088: $f$ has no real zeros at all, while all the zeros of $g$
1089: are contained in $|z'|<2$. Both of these facts are established by Sturm's
1090: criterion. It follows that the maximal value of $z'$ is less that $2$, and this concludes the proof of
1091: Proposition~\ref{prop:nospi}.
1092: \qed
1093:
1094: {\bf Acknowledgment} This research was completed while the second author was a member at the Institute for Advanced Study, Princeton NJ. He wishes to thank the institute and its staff for their hospitality.
1095:
1096: \bibliographystyle{plain}
1097: \def\cprime{$'$}
1098: \begin{thebibliography}{10}
1099:
1100: \bibitem{Bey01}
1101: Horst~R. Beyer.
1102: \newblock On the stability of the {K}err metric.
1103: \newblock {\em Comm. Math. Phys.}, 221(3):659--676, 2001.
1104:
1105: \bibitem{Bic78}
1106: J.~Bi{\v{c}}ak.
1107: \newblock Gravitational collapse with charge and small asymmetries, I. Scalar perturbations.
1108: \newblock {\em Gen. Rel. Grav.} 3:331-349, 1972.
1109:
1110: \bibitem{BluSof03}
1111: P.~Blue and A.~Soffer.
1112: \newblock Semilinear wave equations on the {S}chwarzschild manifold. {I}.
1113: {L}ocal decay estimates.
1114: \newblock {\em Adv. Differential Equations}, 8(5):595--614, 2003.
1115:
1116: \bibitem{BluSof03b}
1117: P.~Blue and A~Soffer.
1118: \newblock The wave equation on the {S}chwarzschild metric {II}: Local decay for
1119: the spin 2 {R}egge-{W}heeler equation.
1120: \newblock arXiv:gr-qc/0310066.
1121:
1122: \bibitem{BPST2}
1123: Nicolas Burq, Fabrice Planchon, John Stalker, and A.~Shadi Tahvildar-Zadeh.
1124: \newblock Strichartz estimates for the wave and {S}chr\"odinger equations with
1125: potentials of critical decay.
1126: \newblock arXiv:math.AP/0401019.
1127:
1128: \bibitem{Cla98}
1129: C.~J.~S.~Clarke.
1130: \newblock Generalized hyperbolicity in singular spacetimes.
1131: \newblock {\em Class. Quantum Grav.}, 15:975--984, 1998.
1132:
1133: \bibitem{CYDL03}
1134: V. Cardoso, S. Yoshida, O. J. C. Dias and J. P. S. Lemos,
1135: \newblock Late-Time Tails of Wave Propagation in Higher Dimensional Spacetimes.
1136: \newblock {\em Phys. Rev. D} 68 (2003) 061503 (Rapid Communications); arXiv:hep-th/0307122.
1137:
1138: \bibitem{DafRod03}
1139: Mihalis Dafermos and Igor Rodnianski.
1140: \newblock A proof of {P}rice s law for the collapse of a self-gravitating
1141: scalar field.
1142: \newblock arXiv:gr-qc/0309115.
1143:
1144: \bibitem{FKSY02}
1145: F.~Finster, N.~Kamran, J.~Smoller, and S.-T. Yau.
1146: \newblock Decay rates and probability estimates for massive {D}irac particles
1147: in the {K}err-{N}ewman black hole geometry.
1148: \newblock {\em Comm. Math. Phys.}, 230(2):201--244, 2002.
1149:
1150: \bibitem{HorMar95}
1151: Gary~T. Horowitz and Donald Marolf.
1152: \newblock Quantum probes of spacetime singularities.
1153: \newblock {\em Phys. Rev. D (3)}, 52(10):5670--5675, 1995.
1154:
1155: \bibitem{IshHos99}
1156: Akihiro Ishibashi and Akio Hosoya.
1157: \newblock Who's afraid of naked singularities? {P}robing timelike singularities
1158: with finite energy waves.
1159: \newblock {\em Phys. Rev. D (3)}, 60(10):104028, 12, 1999.
1160:
1161: \bibitem{IshWal03}
1162: Akihiro Ishibashi and Robert~M. Wald.
1163: \newblock Dynamics in non-globally-hyperbolic static spacetimes. {II}.
1164: {G}eneral analysis of prescriptions for dynamics.
1165: \newblock {\em Classical Quantum Gravity}, 20(16):3815--3826, 2003.
1166:
1167: \bibitem{KayWal87}
1168: Bernard~S. Kay and Robert~M. Wald.
1169: \newblock Linear stability of {S}chwarzschild under perturbations which are
1170: nonvanishing on the bifurcation {$2$}-sphere.
1171: \newblock {\em Classical Quantum Gravity}, 4(4):893--898, 1987.
1172:
1173: \bibitem{KoyTom01a}
1174: Hiroko Koyama and Akira Tomimatsu.
1175: \newblock Asymptotic tails of massive scalar fields in a
1176: {R}eissner-{N}ordstr\"om background.
1177: \newblock {\em Phys. Rev. D (3)}, 63(6):064032, 9, 2001.
1178:
1179: \bibitem{KoyTom01b}
1180: Hiroko Koyama and Akira Tomimatsu.
1181: \newblock Asymptotic tails of massive scalar fields in a {S}chwarzschild
1182: background.
1183: \newblock {\em Phys. Rev. D (3)}, 64(4):044014, 8, 2001.
1184:
1185: \bibitem{MacSta04}
1186: Matei Machedon and John~G. Stalker.
1187: \newblock Decay of solutions to the wave equation on a spherically symmetric
1188: static background.
1189: \newblock {\em in preparation}.
1190:
1191: \bibitem{Mon74}
1192: Vincent Moncrief.
1193: \newblock Odd-parity stability of a {R}eissner-{N}ordstr\"om black hole.
1194: \newblock {\em Phys. Rev. D (3)}, 9:2707, 1974.
1195:
1196: \bibitem{Poi03}
1197: Eric Poisson.
1198: \newblock Radiative falloff of a scalar field in a weakly curved spacetime
1199: without symmetries.
1200: \newblock {\em Phys. Rev. D (3)}, 66(4):044008, 17, 2002.
1201:
1202: \bibitem{Pri72}
1203: Richard Price.
1204: \newblock Nonspherical perturbations of relativistic gravitational collapse,
1205: {I-II}.
1206: \newblock {\em Phys. Rev. D (3)}, 5:2419--2438,2439--2454, 1972.
1207:
1208: \bibitem{ReeSimII}
1209: Michael Reed and Barry Simon.
1210: \newblock {\em Methods of modern mathematical physics. {II}. {F}ourier
1211: analysis, self-adjointness}.
1212: \newblock Academic Press [Harcourt Brace Jovanovich Publishers], New York,
1213: 1975.
1214:
1215: \bibitem{Seg03}
1216: Itai Seggev.
1217: \newblock Dynamics in stationary, non-globally hyperbolic spacetimes.
1218: \newblock arXiv:gr-qc/0310016.
1219:
1220: \bibitem{Sta03b}
1221: John~G. Stalker.
1222: \newblock A compactness criterion for real plane algebraic curves.
1223: \newblock {\em preprint}, 2003.
1224:
1225: \bibitem{Neu30}
1226: John Von~Neumann.
1227: \newblock Allgemeine Eigenwerttheorie {H}ermitescher Funktionaloperatoren.
1228: \newblock {\em Math. Ann.}, 102:49--131, 1929-1930.
1229:
1230: \bibitem{Wal80}
1231: Robert~M. Wald.
1232: \newblock Dynamics in nonglobally hyperbolic, static space-times.
1233: \newblock {\em J. Math. Phys.}, 21(12):2802--2805, 1980.
1234:
1235: \bibitem{Whi89}
1236: Bernard~F. Whiting.
1237: \newblock Mode stability of the {K}err black hole.
1238: \newblock {\em J. Math. Phys.}, 30(6):1301--1305, 1989.
1239:
1240: \end{thebibliography}
1241:
1242: \par\noindent $^{\mathbf a}$
1243: %%%%%%%%%%%%%%%%%%%%%%% Adresse no 1 %%%%%%%%%%%%%%%%%
1244: Department of Mathematics\\
1245: Princeton University,\\
1246: Princeton NJ 08544\\
1247: Email: {\sl stalker@math.princeton.edu}
1248: \ \\
1249: \\ $^{\mathbf b}$
1250: %%%%%%%%%%%%%%%%%%%%%%% Adresse no 2 %%%%%%%%%%%%%%%%%
1251: Department of Mathematics\\
1252: Rutgers, The State University of New Jersey\\
1253: 110 Frelinghuysen Road, Piscataway NJ 08854\\
1254: Email: {\sl shadi@math.rutgers.edu}
1255:
1256:
1257: \end{document}
1258: