gr-qc0401009/CQG.tex
1: %&LaTeX
2: \documentclass[12pt]{iopart}
3: \usepackage[]{indentfirst}
4: \usepackage{amssymb}
5: 
6: 
7: \begin{document}
8: 
9: 
10: 
11: \begin{titlepage}
12: 
13: 
14: \title{Quantum Singularity of Levi-Civita Spacetimes}
15: 
16: \author{D A Konkowski\dag, T M  Helliwell\ddag,   and C Wieland\ddag}
17: 
18: \address{\dag Department of Mathematics, U.S. Naval Academy, Annapolis,
19: Maryland, 21402 U.S.A.}
20: 
21: \address{\ddag Department of Physics, Harvey Mudd College,
22: Claremont, California 91711 USA}
23: 
24: \eads{\mailto{dak@usna.edu}, \mailto{T\_Helliwell@HMC.edu}}
25: 
26: 
27: 
28: \begin{abstract}
29:         Quantum singularities in general relativistic spacetimes are
30:         determined by the behavior of quantum test particles. A static
31:         spacetime is quantum mechanically singular if the spatial
32:         portion of the wave operator is not essentially self-adjoint.
33:         Here Weyl's limit point-limit circle criterion is used to
34:         determine whether a wave operator is essentially self-adjoint.
35:         This test is then applied to scalar wave packets in
36:         Levi-Civita spacetimes to help elucidate the physical
37:         properties of the spacetimes in terms of their metric parameters.
38: \end{abstract}
39: 
40: \pacs{04.20 Dw, O4.62.+v, 03.65 Db}
41: 
42: \maketitle
43: 
44: \end{titlepage}
45: 
46:  \section{Introduction}
47:  In classical general relativity, singularities are not part of the
48: spacetime;
49:  they are boundary points indicated by incomplete geodesics or
50:  incomplete curves of bounded acceleration in a maximal
51:  spacetime (see, e.g. \cite{HE, ES}). So at least for timelike
52:  and null geodesics, this incompleteness can be considered as the
53:  abrupt ending of classical particle paths.
54: 
55:  \par What happens if, instead of classical particles, one uses quantum
56:  mechanical particles to identify singularities? This is the question G.
57:  Horowitz and D. Marolf \cite{HM} set out to answer following earlier
58:  work by R. Wald \cite{Wald}. They call a spacetime quantum
59:  mechanically nonsingular if the evolution of a test scalar wave
60:  packet in the spacetime, representing a quantum mechanical particle,
61:  is uniquely determined by the initial wave
62:  packet, without having to place arbitrary boundary conditions at the
63:  classical singularity. Their construction is restricted to
64:  static spacetimes.
65: 
66:  \par Mathematically, this evolution is related to properties of a
67:  quantum mechanical operator. A static spacetime is quantum mechanically
68:  singular \cite{HM} if the spatial portion of the Klein-Gordon wave
69: operator is not
70:  essentially self-adjoint \cite{RS} \cite{Rich}. If this operator is not
71:  essentially self-adjoint then the evolution of a test scalar wave
72:  packet is not determined uniquely by the initial wave packet;
73:  boundary conditions at the classical singularity are needed to ``pick
74:  out'' the correct wave function and thus one needs to add information
75:  that is not already present in the wave operator, spacetime metric,
76:  and manifold. This will be described further in the paragraphs below.
77: 
78:  \par Horowitz and Marolf restrict the wave operator to the Klein-Gordon
79:  operator, but it has been shown \cite{HKA} that Maxwell and Dirac operators
80:  may also be used with equivalent generic results. Throughout this paper
81:  we will restrict the analysis to Klein-Gordon operators.
82: 
83:  \par A relativistic quantum particle with mass $M$ can be described
84:  by a positive frequency solution to the Klein-Gordon equation
85: 
86:  \begin{equation}
87:      \frac{\partial ^{2} \Psi}{\partial t^{2}} = -A \Psi
88:      \label{eq:1}
89:  \end{equation}
90: 
91: 
92: \noindent in a static spacetime \cite{HM}. The spatial Klein-Gordon
93: operator $A$
94: is
95: 
96: \begin{equation}
97:     A \equiv -V D^{i}(V D_{i}) + V^{2} M^{2}
98:     \label{eq:2}
99: \end{equation}
100: 
101: \noindent where $V^{2} = - \xi_{\nu}\xi^{\nu}$ (here $\xi^{0}$ is the
102: timelike
103: Killing field) and $D_{i}$ is the spatial covariant derivative on the
104: static slice $\Sigma$. The appropriate Hilbert space $H$ is
105: $\mathcal{L}^{2}(\Sigma)$, the space of square integrable functions on
106: $\Sigma$.
107: This choice agrees with that of Horowitz and Marolf \cite{HM}.  The
108: volume element used to define $H$ is $V^{-1}$ times the natural volume
109: element on $\Sigma$. We do
110: not choose the first Sobolev norm $H^{1}$ proposed by Ishibashi and
111: Hosoya \cite{IH}; that choice is related to Dirichlet boundary
112: conditions at the singularity \cite{Wald}. If we initially define
113: the domain of $A$ to be $C_{0}^{\infty}$, $A$ is a real positive
114: symmetric operator and self-adjoint extensions always exist \cite{RS}.
115: If there is one unique self-adjoint extension $A_{E}$, then $A$ is
116: essentially self-adjoint \cite{RS}. In this case the Klein-Gordon
117: equation for a free relativistic particle takes the form \cite{HM}
118: 
119: \begin{equation}
120:     i \frac{\partial \Psi}{\partial t} = (A_{E})^{1/2} \Psi
121:     \label{eq:3}
122: \end{equation}
123: 
124: \noindent with
125: 
126: \begin{equation}
127:    \Psi (t) = \exp ( -i t (A_{E}^{1/2}) \Psi (0).
128:     \label{eq:4}
129: \end{equation}
130: 
131: \noindent Equations (\ref{eq:3}) and (\ref{eq:4}) are ambiguous if $A$ is
132: not essentially
133: self-adjoint. This fact led Horowitz and Marolf \cite{HM} to define
134: quantum-mechanically singular spacetimes as those in which $A$ is not
135: essentially self-adjoint.
136: 
137: \par In their paper Horowitz and Marolf test several classically
138: singular spacetimes to determine whether they are quantum mechanically
139: singular as well. They find that Reissner-Nordstr\"{o}m, negative mass
140: Schwarzschild, and the $2D$ cone remain singular when probed by
141: quantum scalar test particles; however, certain orbifolds, extreme
142: Kaluza-Klein black holes, the $D=5$ fundamental string, and a few other
143: examples are nonsingular. Helliwell and Konkowski \cite{HK} and
144: Helliwell, Konkowski and Arndt \cite{HKA} show that a broad class of
145: quasiregular spacetimes (spacetimes with quasiregular singularities)
146: are quantum mechanically singular. This work was an extension of
147: earlier work by Kay and Studer \cite{KS} which showed that the $2D$ cone and
148: a $4D$ idealized cosmic string are not essentially self-adjoint.
149: 
150: \par The work of Horowitz and Marolf \cite{HM} and subsequent work by
151: Helliwell, Konkowski and Arndt \cite{HKA} test for essential
152: self-adjointness of the spatial
153: operator $A$ by using a von Neumann criterion \cite{VN, weyl} . This method
154: involves studying solutions to
155: 
156: \begin{equation}
157:     A \Psi = \pm i \Psi
158:     \label{eq:5}
159: \end{equation}
160: 
161: \noindent and finding the number of solutions that belong to
162: $\mathcal{L}^{2}(\Sigma)$
163: for each sign of $i$. This determines the von Neumann deficiency indices
164: which indicate whether the operator is essentially self-adjoint or
165: whether it has self-adjoint extensions, and how many self-adjoint
166: extensions it has \footnote{For an
167: introduction to the mathematics of self-adjoint operators and von
168: Neumann deficiency indices, see \cite{HKA} which is based on \cite{RS, IH,
169: BFV}} . If
170: the deficiency indices are $(0,0)$, so that no solutions are square
171: integrable, then the operator is essentially self-adjoint and so has
172: a unique self-adjoint extension. The wave behavior described by the
173: operator is uniquely determined for all time by the operator, the
174: spacetime metric, and the manifold. No additional information in the
175: form of boundary conditions need be added.
176: 
177: \par In this paper we use an alternative test for quantum
178: singularity based upon the limit point-limit circle criterion for
179: essential self-adjointness of operators. That is, we use a theorem of Weyl's
180: \cite{RS} to relate the essential self-adjointness of the operator via
181: the von Neumann deficiency indices to the ``potential'' which
182: determines the behavior of the scalar wave packet. The effect is
183: determined by the limit point-limit circle criterion
184: which we will discuss in Section 3. After developing this technique in
185: the context of determining the quantum singularity nature of a general
186: spacetime, we
187: study the so-called Levi-Civita spacetimes in particular.
188: 
189: \par Levi-Civita spacetimes are static, primarily cylindrically symmetric,
190: spacetimes that are classically singular at ``$r=0$'' unless the
191: metric is Minkowski or Minkowski in accelerated coordinates. They are
192: used to model infinite line masses and idealized cosmic strings.
193: There has been some controversy in the literature about the physical
194: relevance of certain parameters; this will be reviewed in Section 2.
195: The quantum singularity or nonsingularity of $r=0$ will be studied in
196: Section 4 (following the mathematical background in Section 3).  Finally
197: in Section 5
198: we use the quantum singularity results
199: for certain parameter values to glean what information we can about the
200: physical
201: interpretation of the corresponding spacetimes.
202: 
203: \section{Levi-Civita Spacetimes}
204: 
205: The metric in Levi-Civita spacetimes \cite{LC} has the form
206: 
207: \begin{equation}
208:     ds^{2} = r^{4 \sigma}dt^{2} - r^{8 \sigma^{2} - 4\sigma}(dr^{2}+
209:     dz^{2}) - \frac{r^{2 - 4 \sigma}}{C^{2}} d\theta^{2}
210:     \label{eq:6}
211: \end{equation}
212: 
213: \noindent where $\sigma$ and $C$ are real numbers ($C>0$).  It is
214: straightforward to show that if $\sigma$ is replaced by $1/4\sigma$
215: in the metric, the spacetime is unchanged, although a coordinate
216: transformation is required to make this obvious.  Therefore we
217: need only study the range $-\frac{1}{2}\le \sigma\le\frac{1}{2}$.
218: Neither of the parameters $\sigma$ and $C$ can
219: be removed by a coordinate transformation \cite{bonnor,
220: HSTW, HRS}.  Computation of the Kretschmann scalar
221: 
222: \begin{equation}
223:     R_{\mu \nu \sigma \tau}R^{\mu \nu \sigma \tau} = \frac{64
224:     \sigma^{2} (2\sigma -1)^{2}}{(4 \sigma^{2} -2\sigma +1)^{3} r^{4}}
225:     \label{eq:7}
226: \end{equation}
227: 
228: \noindent shows that $r=0$ is a scalar curvature singularity for all
229: $\sigma $
230: except $\sigma = 0, 1/2$.
231: 
232: \par If $\sigma = 0$, the metric is
233: 
234: \begin{equation}
235:     ds^{2} = -dt^{2} + dr^{2} +dz^{2} + \frac{r^{2}}{C^{2}} d \theta^{2}.
236:     \label{eq:8}
237: \end{equation}
238: 
239: \noindent If $C=1$, this is simply Minkowski spacetime in
240: cylindrical coordinates. It thus seems reasonable to choose the
241: coordinate ranges
242: 
243: \begin{equation}
244:     -\infty < t < \infty,\,  -\infty < z < \infty,\,
245:     0 < r < \infty, \,  0  \le \theta \le 2 \pi,
246:     \label{eq:9}
247: \end{equation}
248: 
249: \noindent with $0$ and $2\pi$ identified. If $C \not= 1$, equation
250: \ref{eq:8} is the
251: metric for an idealized cosmic string. There is a quasiregular
252: (``disclination'') singularity at $r=0$ (see, e.g., \cite{HKA} for a
253: discussion). This is a
254: topological singularity, not a curvature singularity. It is thus
255: reasonable to identify the parameter $C$ in the general metric,
256: equation (6), with a topological property of the spacetime, that is,
257: the deficit angle.
258: 
259: \par The sigma parameter is less easily interpreted \cite{bonnor,
260: HSTW, HRS}. For certain values one can interpret these
261: spacetimes as those of infinite line masses. This
262: interpretation is clearest for $|\sigma|\ll 1$. The most physically
263: realistic range is $0<\sigma<1/4$,  where $\sigma$
264: represents the mass per unit length and timelike circular orbits exist.
265: As $\sigma$ increases from 1/4 to 1/2 the interpretation is more
266: difficult \cite{bonnor, HSTW, HRS}. In this range, the Kretschmann
267: scalar decreases with increasing $\sigma$, which seems to suggest that the
268: gravitational field is becoming weaker. But as several authors comment
269: \cite{bonnor, HSTW, HRS}, the Kretschmann scalar may not be a good
270: indicator of gravitational
271: strength; the acceleration of test particles, increasing with
272: increasing $\sigma$ in the interval $(1/4, 1/2)$, may be a better
273: indicator.  Interior solutions matching to exterior Levi-Civita spacetimes
274: exist for the entire range $0<\sigma<1/2$. Bonnor \cite{bonnor}
275: proposes to interpret Levi-Civita as the spacetime generated by a
276: cylinder whose radius increases with increasing $\sigma$ and tends to
277: infinity as $\sigma$ tends to $1/2$. Thus the entire range
278: $0<\sigma<1/2$ can be taken to represent an ``infinite line mass''.
279: 
280: 
281: \par The $\sigma = 1/2$ case is more problematic. Bonnor's analysis
282: suggests (but does not prove) that when $\sigma = 1/2$ the cylinder
283: becomes a plane, an interpretation first put forward by Gautreau and
284: Hoffmann \cite{GH} based on different considerations. In fact, the $\sigma
285: =1/2$ metric
286: 
287: \begin{equation}
288:     ds^{2} = -r^{2} dt^{2} + dr^{2} + dz^{2} + \frac{1}{C^{2}}
289:     d\theta^{2}
290:     \label{eq:10}
291: \end{equation}
292: 
293: \noindent is flat. We can transform to Minkowski coordinates
294: 
295: \begin{equation}
296:     ds^{2} = -d{\bar t}^{2} + d{\bar x}^{2} + d{\bar y}^{2} + d{\bar
297:     z}^{2}
298:     \label{eq:11}
299: \end{equation}
300: 
301: \noindent if we let
302: ${\bar t} = r \sinh t$, ${\bar x} = r \cosh t$, ${\bar y} = \theta/C$,
303: and ${\bar z} = z$. Here we can obviously let the ${\bar y}$ coordinate range
304: from $-\infty$ to $\infty$. This is flat spacetime described from the
305: view of an accelerating frame of reference. This seems to support the
306: interpretation of the $\sigma = 1/2$ case as a planar source producing
307: flat spacetime described by a uniformly accelerating observer
308: \cite{bonnor}. In other words, one can interpret it as the spacetime of a
309: gravitational field produced by an
310: infinite planar sheet of positive mass density.
311: 
312: \par On the other hand, one can interpret the metric with $\sigma = -1/2$
313: as the
314: gravitational field produced by an infinite sheet of negative mass
315: density (an idea also first proposed by Gautreau and Hoffmann
316: \cite{GH}). However, this case is also problematic \cite{bonnor, HSTW,
317: HRS}. The metric
318: 
319: \begin{equation}
320:     ds^{2} = - \frac{1}{r^{2}} dt^{2} + r^{2} (dr^{2} + dz^{2} +
321:     \frac{1}{C^{2}} d \theta^{2})
322:        \label{eq:12}
323: \end{equation}
324: 
325: \noindent is not flat and it has a curvature singularity at $r=0$. It
326: has planar symmetry with one additional extra Killing vector beyond
327: those in the general
328: Levi-Civita spacetimes. The interpretation of this
329: spacetime as one caused by an infinite plane of negative mass density
330: comes from the fact that test particles are repelled from the $r=0$
331: plane and the fact that ``the Gaussian curvature of spacelike
332: `eigensurfaces' is
333: zero'' \cite{GH}. This is a reasonable interpretation; the
334: ``problem'' comes from interpreting both $\sigma = 1/2$ and $\sigma
335: =-1/2$ as planes of infinite (positive/negative) mass density as one
336: is singularity-free and the other has a curvature singularity
337: \cite{HSTW, HRS}.
338: 
339: \par We will now study the quantum singularity properties of these
340: spacetimes for various parameter values in the following sections. We
341: begin by introducing the mathematics necessary to discuss essentially
342: self-adjoint operators and quantum singularities.
343: 
344: \section{Weyl's Limit Point - Limit Circle Criterion}
345: 
346: \par  A particularly convenient way to establish essential
347: self-adjointness in the spatial operator of the Klein-Gordon equation
348: is to use the concepts of limit circle and limit point behavior.\footnote
349: {This section is based on {\bf Appendix to X.1} in Reed and
350: Simon \cite{RS}}  The
351: approach is as follows.  The Klein-Gordon equation for Levi-Civita
352: spacetimes can be separated in the coordinates $t, r, \theta, z$.
353: Only the radial equation is non-trivial.  With changes in both
354: dependent and independent variables, the radial equation can be
355: written as a one-dimensional Schr\"{o}dinger equation
356: 
357: \begin{equation}
358:     H\Psi(x) = E\Psi(x)
359:     \label{eq:13}
360: \end{equation}
361: 
362: \noindent where $x \in (0,\infty )$ and the operator  $H = - d^{2}/dx^{2}
363: + V(x)$.
364: 
365: \newtheorem{Definition}{Definition}
366: \begin{Definition} The potential $V(x)$ is in the limit circle case
367: at $x = 0$  if for some, and therefore for all $E$, {\it all}
368: solutions of equation (\ref{eq:13}) are square integrable at zero.  If
369: $V(x)$  is not
370: in the limit circle case, it is in the limit point case.
371: \end{Definition}
372: 
373: \par  A similar definition pertains to $x=\infty$.  The potential
374: $V(x)$ is in the limit circle case at $x=\infty$ if all solutions of
375: equation (\ref{eq:13}) are square integrable at infinity; otherwise,
376: $V(x)$ is in
377: the limit point case at infinity.
378: 
379: \par There are of course two linearly independent solutions of the
380: Schr\"{o}dinger equation for given $E$.  If $V(x)$ is in the limit circle
381: case at zero, both solutions are $\mathcal{L}^{2}$ at zero, so all linear
382: combinations are $\mathcal{L}^{2}$ as well.  We would therefore need a
383: boundary
384: condition at $x=0$ to establish a unique solution.  If $V(x)$ is in
385: the limit {\it point} case, the $\mathcal{L}^{2}$ requirement eliminates
386: one of
387: the solutions, leaving a unique solution without the need of
388: establishing a boundary condition at $x=0$. This is the whole idea of
389: testing for quantum singularities; there is no singularity if the
390: solution is unique, as it is in the limit point case.  The critical
391: theorem is due to Weyl \cite{RS}.
392: 
393: \newtheorem{theorem}{Theorem}
394: \begin{theorem}[The Weyl limit point-limit circle criterion.] If
395: $V(x)$ is a continuous real-valued function on $(0, \infty)$, then
396: $H = - d^{2}/dx^{2} + V(x)$ is essentially self-adjoint on
397: $C_{0}^{\infty}(0, \infty)$ if and only if $V(x)$ is in the limit
398: point case at both zero and infinity.
399: \end{theorem}
400: 
401: \par The following theorem can be used to establish the limit circle-limit
402: point behavior at infinity \cite{RS}.
403: 
404: \newtheorem{theorem1}[theorem]{Theorem}
405: \begin{theorem1}[Theorem X.8 of Reed and Simon \cite{RS}.] If $V(x)$ is
406: continuous and real-valued on $(0, \infty)$, then $V(x)$ is in the
407: limit point case at infinity if there exists a {\em positive}
408: differentiable function $M(x)$ so that
409: \begin{enumerate}
410:     \item[(i)] $V(x) \ge - M(x)$
411:     \item[(ii)] $\int_{1}^{\infty} [M(x)]^{-1/2} dx = \infty$
412:     \item[(iii)] $M'(x)/M^{3/2}(x)$ is bounded near $\infty$.
413: \end{enumerate}
414: Then $V(x)$ is in the limit point case (complete) at $\infty$.
415: \end{theorem1}
416: 
417: A sufficient choice of the $M(x)$ function for our purposes is the
418: power law function $M(x) = c x^{2}$ where $c > 0$. Then {\it (ii)} and
419: {\it (iii)} are satisfied, so if $V(x) \ge -c x^{2}$, $V(x)$ is in the
420: limit point case at infinity.
421: 
422: \par A theorem useful near zero is the following.
423: 
424: \newtheorem{theorem2}[theorem]{Theorem}
425: \begin{theorem2} [Theorem X.10 of Reed and Simon \cite{RS}.]
426: Let $V(x)$ be continuous and {\it positive} near zero.
427: If $V(x)\ge\frac{3}{4} x^{-2}$ near zero then $V(x)$ is in the limit
428: point case.  If for some $\epsilon > 0$,
429: $V(x)\le(\frac{3}{4}-\epsilon)x^{-2}$ near zero, then $V(x)$ is in the
430: limit circle case.
431: \end{theorem2}
432: 
433: \noindent These results can now be used to help test for quantum
434: singularities in the Levi-Civita spacetimes.
435: 
436: \section{Limit Point - Limit Circle properties of Levi-Civita Spacetimes}
437: 
438: \par The Klein-Gordon equation for a scalar particle of mass $M$ is
439: 
440: \begin{equation}
441:      \Box \Phi \equiv g^{\mu \nu} \Phi_{\,\mu \nu} +
442:      \frac{1}{\sqrt{g}}(\sqrt{g}g^{\mu \nu}),_{\nu} \Phi_{\,\mu} = M^{2}
443:      \Phi ;
444:      \label{eq:14}
445: \end{equation}
446: 
447: \noindent for the Levi-Civita metrics it is
448: 
449: \begin{equation}
450:     \fl -r^{-4 \sigma} \Phi,_{tt} + r^{8 \sigma^{2} + 4 \sigma}
451:     (\Phi,_{rr} + \Phi,_{zz}) +C^{2} r^{4 \sigma - 2}
452:     \Phi,_{\theta\theta} +
453:     r^{-8 \sigma^{2} + 4\sigma - 1} \Phi,_{r} = M^{2} \Phi.
454:     \label{eq:15}
455: \end{equation}
456: 
457: \noindent The equation separates using $\Phi = e^{-i \omega
458: t}e^{ikz}e^{im \theta} R(r)$, where $\omega, k$ are continuous and $m=0,
459: \pm 1, \pm 2, \cdots$. The resulting radial equation is
460: 
461: \begin{equation}
462:     \frac {d^{2} R}{d r^{2}} + \frac{1}{r} \frac{d R}{d r} + \left[
463:     \omega^{2} r^{8 \sigma^{2}- 8 \sigma} - k^{2} - M^{2} r^{8
464:     \sigma^{2} - 4 \sigma}-m^{2} C^{2} r^{8 \sigma^{2} - 2} \right]
465:     R = 0.
466:     \label{eq:16}
467: \end{equation}
468: 
469: \noindent As described in Section 1, square integrability is judged by
470: evaluating
471: the integral
472: 
473: \begin{equation}
474:     I = \int dr \sqrt{g_{3}/g_{00}} R^{*} R = (1/C) \int dr  r^{8
475:     \sigma^{2} - 8 \sigma +1} R^{*} R
476:     \label{eq:17}
477: \end{equation}
478: 
479: \noindent where $g_{3}$ is the negative determinant of the 3-space portion of
480: the metric tensor \cite{Wald, HM}.  The resulting integral is
481: invariant under coordinate transformations.
482: 
483: \par To test for limit point-limit circle behavior one must
484: rewrite the radial equation in the one-dimensional
485: Schr\"{o}dinger-equation form
486: 
487: \begin{equation}
488:     \frac{d^{2} \Psi (x)}{dx^{2}} + \left[E - V(x) \right] \Psi = 0,
489:     \label{eq:18}
490: \end{equation}
491: 
492: \noindent for which square integrability is judged by the integral $I =
493: \int dx
494: \Psi^{*} \Psi$.  Except for the special case $\sigma = 1/2$, which we
495: return to later, the
496: appropriate transformations of dependent and independent variables are
497: $R = (1/\sqrt{x}) \Psi$ and $r = ( \alpha C x^{2})^{1/(2 \alpha)}$
498: where $\alpha \equiv (2 \sigma - 1)^{2}$.
499: 
500: \par Equation (16) transforms to equation (18) if we let $E = C
501: \omega^{2}/ \alpha$ and
502: \begin{eqnarray}
503:     V(x) & = & (C k^{2}/ \alpha) ( \alpha C x^{2})^{(- \alpha + 1)/
504:     \alpha} + (C M^{2}/ \alpha) ( \alpha C x^{2})^{2 \sigma/
505:     \alpha} \nonumber\\
506:       & + & (m^{2} C^{3}/ \alpha) (\alpha C x^{2})^{(4 \sigma - 1)/
507:     \alpha} - 1/(4 x^{2}).
508:     \label{eq:19}
509: \end{eqnarray}
510: 
511: \noindent In the limit $x \rightarrow \infty$ the final term vanishes,
512: leaving
513: a non-negative $V(x)$. Therefore the solutions are limit point at
514: infinity for all $\sigma$ ($\sigma \not= 1/2$), since as $x \rightarrow
515: \infty$, $V(x) > - c x^{2}$ for any positive constant $c$ \cite{RS,
516: Rich}.
517: 
518: \par It remains to test the solutions in the limit $x \rightarrow 0$.
519: This we do by identifying the important terms in $V(x)$. The final term
520: diverges as $x^{-2}$. The first term in $V(x)$ diverges less quickly for
521: any finite  $\sigma$. The second term $\sim x^{4 \sigma/ \alpha}$,
522: whose exponent has the minimum value $-1/2$ at $\sigma = -1/2$. The
523: third term $\sim x^{2( 4 \sigma - 1)/ \alpha}$, whose exponent has the
524: minimum value -2 at $\sigma = 0$.
525: 
526: \par Therefore unless $\sigma = 0$ (or $\sigma = 1/2$, which we
527: treat later), the equation has the form
528: 
529: \begin{equation}
530:     \frac{d^{2} \Psi}{dx^{2}} + \frac{1}{4 x^{2}} \Psi = 0
531:     \label{eq:20}
532: \end{equation}
533: 
534: \noindent as $x \rightarrow 0$.  By Theorem 3 the solutions
535: should exhibit limit circle behavior; in fact, one solution is $\Psi_{1} =
536: x^{1/2}$
537: and the other is $\Psi_{2} = \Psi_{1} \int^{x} dx/
538: \Psi_{1}^{2} = x^{1/2} \ln (x)$. Both are $\mathcal{L}^{2}$
539: near $x = 0$, so are indeed limit circle.
540: 
541: \par If $\sigma =0$, the equation has the form
542: \begin{equation}
543:     \frac{d^{2} \Psi}{d x^{2}} - \frac{m^{2} C^{2} - 1/4}{x^{2}}
544:     \Psi = 0
545:     \label{eq:21}
546: \end{equation}
547: 
548: \noindent as $x \rightarrow 0$. One solution is $\Psi_{1} = x^{1/2 + |m| C}$;
549: the other is $\Psi_{2} = x^{1/2} \ln (x)$ (if $m = 0$), or $
550: \Psi_{2} = x^{1/2 - |m| C}$ (if $m \not= 0$). Both solutions are
551: limit circle if $|m| C < 1$, and limit point if $|m| C \ge 1$.
552: 
553: \par In the special case $\sigma = 0,\, C = 1$, the spacetime is
554: Minkowski with no classical singularity at $r =0$. In that case $x =
555: r$ and $R_{1} = 1, R_{2} = \ln (r)$. The solution $R_{2} = \ln (r)$ is
556: unacceptable in this case because it diverges at $r =0$ even though $r
557: =0$ is a
558: regular hypersurface in the spacetime.  \footnote {We assumed at the
559: outset that
560: the domain of the spatial operator consists
561: of smooth functions with compact support away from the origin.
562: For Minkowski space the points $r = 0$ are arbitrary, so it is
563: unphysical to permit a function that diverges at  $r = 0$  and nowhere else.}
564: If $\sigma = 0,\, C \neq 1$ the spacetime
565: corresponds to a linear idealized cosmic string with a quasiregular
566: singularity at $r = 0$.
567: 
568: \par For the special case $\sigma = 1/2$ the radial Klein-Gordon
569: equation is
570: 
571: \begin{equation}
572:     \frac{d^{2} R}{dr^{2}} + \frac{1}{r} \frac{d R}{dr} +\left[ -(
573:     k^{2} + M^{2} + m^{2} C^{2}) + \omega^{2}/ r^{2} \right] R = 0.
574:     \label{eq:22}
575: \end{equation}
576: 
577: \noindent The transformation $R(r) = \Psi (x)$ and $r = \exp (C x)$ puts the
578: equation in the form
579: 
580: \begin{equation}
581:     \frac{d^{2} \Psi}{dx^{2}} + \left[ E - V(x) \right] \Psi = 0
582:     \label{eq:23}
583: \end{equation}
584: 
585: \noindent where $E = C^{2} \omega^{2}$ and
586: 
587: \begin{equation}
588:     V(x) = C^{2} ( k^{2} + M^{2} + m^{2} C^{2}) \exp (2Cx)
589:     \label{eq:24}
590: \end{equation}
591: 
592: \noindent with  $\int dr \sqrt{g_{3}/g_{00}} R^{*} R = \int dx \Psi^{*} \Psi$.
593: 
594: \par In the transformation $r = 0$ corresponds to $x = - \infty$, while
595: $r = \infty $ corresponds to $x = \infty$. So in this case we must
596: determine the limit point-limit circle behavior at $x = \pm \infty $.
597: In this case $V(x) > -c x^{2}$ at both $\pm \infty $, for any
598: positive $c$. Therefore the solutions are limit point at both ends.
599: It is interesting that the $\sigma = 1/2$ case is a special case
600: which must be studied on the interval $(-\infty, \infty)$.  This result seems
601: to support Bonnor's description of the $\sigma = 1/2$ case as a
602: cylinder whose radius tends to $\infty$ as $\sigma$ tends to $1/2$.
603: 
604: 
605: \section{Conclusions}
606: 
607: \par In the preceding section we showed that the Klein-Gordon operator
608: is limit point at infinity for the whole class of Levi-Civita
609: spacetimes. If $\sigma$ is neither zero nor one-half, the
610: Klein-Gordon operator is limit circle at $r = 0$ and thus not
611: essentially self-adjoint,  so all $\sigma \neq
612: 0, \sigma \neq 1/2$ Levi-Civita spacetimes are quantum mechanically
613: singular as well as being classically singular.
614: 
615: \par If $\sigma = 0$ and $C = 1$, the spacetime is simply Minkowski
616: space. One of the two solutions of the radial Klein-Gordon equation
617: can be rejected because it diverges at a regular point ($r=0$) of the
618: spacetime. The operator is therefore limit point at $r=0$, and
619: so Minkowski spacetime is quantum
620: mechanically nonsingular (a well known fact, repeated here for
621: completeness).
622: 
623: 
624: par If $\sigma = 0$ and $C \neq 1$, the spacetime is the conical
625: spacetime corresponding to an idealized cosmic string. The operator
626: is limit circle if $|m| C < 1$ and limit point if $|m| C \ge 1$.
627: Therefore the cosmic string spacetimes are quantum mechanically
628: singular for azimuthal quantum number $m$ such that $|m| C < 1$ and
629: nonsingular if  $|m| C \ge 1$. If arbitrary values of $m$ are allowed,
630: these
631: spacetimes are quantum mechanically singular in agreement with earlier
632: results \cite {HK, HKA, KS}. Recall that these spacetimes are also
633: classically singular with a quasiregular (``disclination'')
634: singularity at $r=0$.
635: 
636: \par If $\sigma = 1/2$ the classical spacetime is flat and without a
637: classical singularity. We have shown that this spacetime is quantum
638: mechanically nonsingular as well for massive scalar particles obeying the
639: Klein-Gordon equation, since the operator is limit point at both ends
640: and thus essentially self-adjoint.
641: 
642: 
643: \par Thus for the Levi-Civita spacetimes, all that are classically
644: singular are also quantum mechanically singular, and all that are
645: classically nonsingular ($\sigma = 0,\, C = 1$, and $\sigma = 1/2$) are
646: also quantum mechanically nonsingular.  The classically and
647: quantum-mechanically nonsingular spacetimes correspond to isolated
648: values of $\sigma$, so that (for example) even though the spacetime
649: $\sigma = 0,\, C = 1$ is nonsingular, the spacetimes with  $\sigma
650: \to 0,\, C = 1$ are singular.  The only discrepency between
651: classical and quantum singularities are for the $\sigma =0,\, C \neq 1$
652: modes with $|m| C \ge 1$, which are quantum mechanically nonsingular
653: in a classically singular spacetime. The physical reason is that the
654: wavefunction for large values of $m$ in a flat space with a quasiregular
655: singularity at $r=0$ are unable to detect the presence of the
656: singularity because of the repulsive centrifugal potential
657: $m^{2}C^{2}/x^{2}$ in equation (21).  The limit point/limit circle
658: criterion, together with the theorems that connect it to the effective
659: potential in a Schrodinger equation, provide physical insight into
660: when quantum singularities are prevented from occurring by potential
661: barriers.
662: 
663: 
664: \ack We thank David Clarke, Jack Harrison and Cassidi Reese for useful
665: conversations.
666: One of us (DAK) was partially funded by NSF grant PHY99-88607 to
667: the U.S. Naval Academy. She also thanks Queen Mary, University of London,
668: where some of this work was carried out.
669: 
670: \section*{References}
671: 
672: \begin{thebibliography}{99}
673: 
674:     \bibitem{HE} Hawking S W and Ellis G F R 1973 {\it The Large-Scale
675:     Structure of Spacetime} (Cambridge: Cambridge University Press)
676: 
677:     \bibitem{ES} Ellis G F R and Schmidt B G 1977 {\it Gen. Relativ. Grav}
678:     {\bf 8} 915
679: 
680:     \bibitem{HM} Horowitz G T and Marolf D 1995 {\it Phys. Rev.
681:     D} {\bf 52} 5670
682: 
683:     \bibitem{Wald} Wald R M 1980 {\it J. Math Phys.} {\bf 21} 2802
684: 
685:     \bibitem{RS} Reed M and Simon B 1972 {\it Functional Analysis}
686:     (New York: Academic Press)
687: 
688:     \nonum Reed M and Simon B 1972 {\it Fourier Analysis and
689:     Self-Adjointness} (New York: Academic Press)
690: 
691:     \bibitem{Rich} Richmyer R -D 1978 {\it Principles of Advanced
692:     Mathematical Physics, Vol. I} (New York: Springer)
693: 
694:     \bibitem{HKA} Helliwell T M, Konkowski D A and Arndt V 2003 {\it
695:     Gen. Relativ. Grav.} {\bf 35} 79
696: 
697:     \bibitem{IH} Ishibashi A and Hosoya A 1999 {\it Phys. Rev. D} {\bf
698:     60} 104028
699: 
700:     \bibitem{HK} Konkowski D A and Helliwell T M 2001 {\it Gen.
701:     Relativ. Grav.} {\bf 33} 1131
702: 
703:     \bibitem{KS} Kay B S and Studer U M 1991 {\it Commun. Math. Phys.}
704:     {\bf 139} 103
705: 
706:     \bibitem{VN} von Neumann J 1929 {\it Math. Ann.} {\bf 102} 49
707: 
708:     \bibitem{weyl} Weyl H  1910 {\it Math. Ann.} {\bf 68} 220
709: 
710:     \bibitem{BFV} Bonneau G, Faraut J and Valent G 2001 {\it Am. J.
711:     Phys.} {\bf 69} 322
712: 
713:     \bibitem{LC} Levi-Civita T 1919 {\it Rend. Acc. Lincei} {\bf 28}
714:     101
715: 
716:     \bibitem{bonnor} Bonnor W B ``The static cylinder in general relativity''
717:     in {\it On Einstein's Path}  ed. A. Harvey 1999 (New York:
718:     Springer) 113
719: 
720:     \bibitem{HRS} Herrera L, Ruifern\'{a}ndez J and Santos N O 2001 {\it Gen.
721:     Relativ. Grav.} {\bf 33} 515
722: 
723:     \bibitem{HSTW} Herrera L, Santos N O, Teixeira A F F, and Wang A Z
724:     2001 {\it Class. Quantum Grav.} {\bf 18} 3847
725: 
726:     \bibitem{GH} Gautreau R and Hoffmann R B 1969 {\it Nuovo Cimento B}
727:     {\bf 61} 411
728: 
729: 
730: 
731: 
732: 
733: \end{thebibliography}
734: 
735: 
736: 
737: \end{document}
738: 
739: 
740: 
741: