1: % PACS number: 0460P
2: % Keywords: Quantum Geometry, Loop Quantum Cosmology
3: %\documentstyle[psfrag,preprint,graphicx,tighten,eqsecnum,floats,aps,amssymb]{revtex}
4: %\documentstyle[psfrag,preprint,graphicx,tighten,floats,aps]{revtex}
5: %\documentclass[preprint,showpacs,preprintnumbers,amsmath,amssymb]{revtex4}
6: \documentclass[preprint,showpacs,preprintnumbers,amsmath,amssymb,nofootinbib]{revtex4}
7: \usepackage{amssymb}
8: \usepackage[mathscr]{eucal}
9: \usepackage{graphicx}
10: \def\be{\begin{equation}}
11: \def\ee{\end{equation}}
12: \def\ba{\begin{eqnarray}}
13: \def\ea{\end{eqnarray}}
14:
15: \def\C{{\cal C}}
16: \def\M{{\cal M}}
17: \def\Gauss{{\cal C}_{\rm Gauss}}
18: \def\ham{{\cal C}_{\rm Ham}}
19: \def\H{{\cal H}}
20: \def\Ab{{\bar{{\cal A}}_{\rm KS}}}
21: \def\cyl{{\rm Cyl_{\rm KS}}}
22: \def\cylstar{{\rm Cyl}^\star_{\rm KS}}
23: \def\tr{{\rm Tr\,}}
24: \def\Tr{{\rm Tr\,}}
25: \def\su{{\rm su}}
26: \def\SU{{\rm SU}}
27:
28: \def\lp{{\ell}_{\rm Pl}}
29: \def\q{{}^o\!q}
30: \def\e{{}^o\!e}
31: \def\w{{}^o\!\omega}
32: \def\L{\delta}
33: \def\f{\frac}
34: \def\T{T}
35: \def\nu{\tau}
36:
37: \newcommand{\md}{{\mathrm d}}
38: \newcommand{\vt}{\vartheta}
39: \newcommand{\vp}{\varphi}
40: \newcommand{\R}{{\mathbb R}}
41: \newcommand{\sgn}{{\mathrm{sgn}}}
42: \def\S{{\mathbb S}}
43: \def\F{{}^o\!F}
44:
45: %\newcommand{\tr}{{\mathrm{tr}}}
46: %\newcommand{\diag}{{\mathrm{diag}}}
47: %\newcommand{\case}[2]{{\textstyle \frac{#1}{#2}}}
48: %\newcommand{\lP}{\ell_{\rm P}}
49: %\newcommand{\rip}[2]{\ensuremath{(#1\,|\,#2\rangle}}
50: %\newcommand{\norm}[1]{\ensuremath{\|#1\|}}
51: %\newcommand{\bra}[1]{\ensuremath{\langle#1|}}
52: %\newcommand{\ket}[1]{\ensuremath{|#1\rangle}}
53: %\newcommand{\ip}[2]{{\langle#1\,|\,#2\rangle}}
54:
55:
56:
57: \begin{document}
58: %\date\today
59:
60: \preprint{IGPG--05--09/01,AEI--2005--132}
61: \title{Quantum geometry and the Schwarzschild singularity}
62: \author{Abhay\ Ashtekar${}^{1,2}$ and Martin Bojowald${}^{2,1}$}
63: \address{1. Institute for Gravitational Physics and Geometry,\\
64: Physics Department, Penn State, University Park, PA 16802, USA\\
65: 2. Max-Planck-Institut f\"ur Gravitationsphysik, Albert-Einstein-Institut,
66: Am M\"uhlenberg 1, D-14476 Potsdam, Germany}
67:
68:
69: \begin{abstract}
70:
71: In homogeneous cosmologies, quantum geometry effects lead to a
72: resolution of the classical singularity without having to invoke
73: special boundary conditions at the singularity or introduce ad-hoc
74: elements such as unphysical matter. The same effects are shown to
75: lead to a resolution of the Schwarzschild singularity. The
76: resulting quantum extension of space-time is likely to have
77: significant implications to the black hole evaporation process.
78: Similarities and differences with the situation in quantum
79: geometrodynamics are pointed out.
80:
81: \end{abstract}
82:
83:
84: \pacs{04.60.Pp, 04.70.Dy, 04.60.Nc, 03.65.Sq}
85: \maketitle
86:
87:
88: \section{Introduction}
89: \label{s1}
90:
91: General relativity provides a subtle and powerful interplay
92: between gravity and geometry, thereby opening numerous
93: possibilities for novel phenomena. Among the most spectacular of
94: the resulting conceptual advances are the predictions that the
95: universe began with a big bang and massive stars can end their
96: lives as black holes. In both cases, one encounters singularities.
97: Space-time of general relativity literally ends and classical
98: physics comes to a halt.
99:
100: However, general relativity is incomplete because it ignores
101: quantum effects. It is widely believed that quantum gravity
102: effects become significant in the high curvature regions that
103: develop {before} singularities are formed. These are likely to
104: significantly change the space-time structure, making the
105: predictions of general relativity unreliable. Hence, real physics
106: need not stop at the big bang and black hole singularities. While
107: the classical space-time does end there, quantum space-time may
108: well extend beyond.
109:
110: Attempts at extending physics beyond the big-bang singularity date
111: back at least to the seventies when mini-superspaces were
112: introduced and quantum cosmology was born (see, e.g., \cite{cm}).
113: These investigations sparked off new developments in different
114: directions ranging from foundations of quantum mechanics to the
115: development of WKB methods to test semi-classicality in quantum
116: cosmology, to the introduction of novel Euclidean methods to
117: calculate the appropriate wave function of the universe. This
118: deepened our understanding of issues related to the quantum
119: physics of the universe as a whole \cite{jh-lh}. However, in these
120: approaches the singularity was not generically resolved. A quantum
121: extension of space-time either required the introduction of new
122: principles \cite{hh}, or ad-hoc assumptions, such as existence of
123: matter violating energy conditions already at the classical level,
124: or of external clocks that remain insensitive to the infinite
125: curvature encountered at singularities.
126:
127: Emergence of quantum Riemannian geometry and the associated
128: mathematical techniques \cite{alrev,crbook,ttrev} have provided a
129: new approach to revisit the issue of quantum extensions. Within
130: the mini-superspaces used in quantum cosmology, the big-bang
131: singularity could be resolved \emph{without} having to introduce
132: external boundary conditions at the big-bang, or matter/clocks
133: with unphysical properties \cite{mb3,mb6,abl}. In the homogeneous,
134: isotropic mini-superspace coupled to a massless scalar field, in
135: particular, an exhaustive analysis can be carried out \cite{aps}.
136: At the analytic level, one can introduce an appropriate inner
137: product on physical states, define Dirac observables and, using
138: the two, construct semi-classical states. One can then use
139: numerical methods to examine the nature of the quantum space-time.
140: If one begins with a semi-classical state at large late times
141: (`now') and evolves it back in time, it remains semi-classical
142: till one encounters the deep Planck regime near the classical
143: singularity. In this regime the quantum geometry effects dominate.
144: However, the state becomes semi-classical again on the other side;
145: the deep Planck region serves as a quantum bridge between two
146: large, classical space-times. The model is too simple to be
147: applied reliably to the actual universe we live in. However at a
148: conceptual and mathematical level it demonstrates the new
149: possibilities, systematically made available by quantum geometry.
150:
151: It is then natural to ask if a similar resolution of singularities
152: occurs also in the context of black holes. Now, there is an
153: extensive literature on the nature of black hole singularities in
154: the classical theory. In particular, detailed mathematical
155: analysis of spherical collapse of uncharged matter was carried out
156: by Christodoulou, Dafermos and others. It has led to the
157: expectation that, for black holes formed through gravitational
158: collapse, the singularity would be generically space-like (see,
159: e.g., \cite{md} and references therein). Therefore, it is natural
160: to focus on this case first and ask for the nature of quantum
161: space-time that would result via their resolution.%
162: %
163: \footnote{Indeed, one does not expect quantum gravity effects to
164: resolve all singularities. If, for example, they led to a
165: resolution of the time-like singularity of the negative energy
166: Schwarzschild space-time, energy would be unbounded below in
167: quantum gravity and the theory would have unphysical features
168: \cite{hm}. Rather, one expects that there would be \emph{no}
169: physical states in quantum gravity that would resemble negative
170: energy Schwarschild geometry at large distances, whence the issue
171: of `resolution' would simply not arise.}
172: %
173: If the singularity is resolved and {if} the quantum geometry in
174: the deep Planck regime again serves as a bridge to a large
175: classical region beyond, there would be no information loss in the
176: black hole formation and evaporation process \cite{ab2} (see also
177: \cite{s'thw,sh3}). Pure states in the distant past would evolve to
178: pure states in the distant future and one would have a space-time
179: description of the entire process in a quantum mechanical setting.
180: Thus, a detailed analysis of the fate of black hole singularities
181: is of considerable importance also for the fundamental issue of
182: whether the standard unitary evolution of quantum physics has to
183: be modified in the setting of black holes.
184:
185: The purpose of this note is to initiate this investigation using
186: quantum geometry, in the setting of connection dynamics. We will
187: focus just on the mini-superspace that is appropriate for
188: describing the geometry interior to the horizon of a Schwarzschild
189: black hole. Since this mini-superspace is spatially homogeneous
190: ---of Kantowski-Sachs type--- one can take over the techniques
191: that have been developed in the setting of homogeneous cosmologies
192: \cite{mb6}. Our main result is that the quantum scalar constraint
193: is indeed such that the singularity is resolved. The salient
194: features of our analysis are as follows. First, we use a
195: \emph{self-adjoint} Hamiltonian constraint. Therefore, our
196: analysis can serve as the point of departure for the construction
197: of the physical Hilbert space either via the group averaging
198: procedure \cite{group} \emph{or} via deparametrization of the
199: theory as in \cite{aps}. Second, we spell out the symmetry
200: reduction procedure that leads to the Kantowski-Sachs
201: mini-superspace \emph{within connection dynamics}. We will see in
202: sections \ref{s3} and \ref{s4} that the structure of the resulting
203: phase space is important to the issue of singularity resolution.%
204: %
205: \footnote{A quantization of the Kantowski-Sachs minisuperspace is
206: available in the geometrodynamical framework, \cite{m}.
207: Unfortunately, it does not shed light on singularity resolution
208: because the location of the singularity in the minisuperspace was
209: misidentified. In the notation used in \cite{m}, the singularity
210: lies at $(a=\infty, b=0)$ ---not at $(a=0,b=0)$ as assumed there.
211: Therefore, neither is the inverse volume operator (17) of \cite{m}
212: bounded or well-defined at the singularity nor does the discrete
213: equation (23) of \cite{m} enable one to evolve across the
214: singularity.}
215: %
216: In particular, this space is an extension of the phase space used
217: in geometrodynamics in that one allows for degenerate triads (and
218: hence 3-metrics). Thanks to this enlargement, points of the
219: reduced phase space corresponding to the singularity do not
220: constitute a boundary, whence the support of the wave function
221: `beyond singularity' can be interpreted geometrically. However, we
222: emphasize that the results of this note constitute only initial
223: steps for a more complete theory, e.g., along the lines of
224: \cite{aps}.
225:
226: Our discussion is organized as follows. In section \ref{s2} we
227: discuss the classical theory of the Kantowski-Sachs
228: mini-superspace using connection-dynamics. In section \ref{s3} we
229: analyze the kinematics of its quantum theory and in section
230: \ref{s4} we present quantum dynamics. Section \ref{s5} provides a
231: brief summary and directions for future work.
232:
233: \section{Classical Theory}
234: \label{s2}
235:
236: This section is divided into two parts. In the first, we carry out
237: the Kantowski-Sachs symmetry reduction of vacuum general
238: relativity and in the second we discuss the structure of the
239: resulting phase space.
240:
241: \subsection{Symmetry reduction}
242: \label{s2.1}
243:
244: The portion of the space-time interior to the horizon of a
245: Schwarzschild black hole can be naturally foliated by 3-manifolds
246: $M$ with topology $\R\times \S^2$ such that the geometry on each
247: slice is invariant under the Kantowski--Sachs symmetry group
248: $G=\R\times {\rm SO(3)}$. In the discussion of this model we
249: closely follow \cite{abl} where the mathematical structure of the
250: simpler, isotropic model is studied in detail.
251:
252: As in any homogeneous situation, $M$ acquires an equivalence class
253: of positive definite metrics related by an overall constant. We
254: fix one such metric $\q_{ab}$ as well as an orthonormal triad
255: $\e_i^a$ and co-triad $\w_a^i$, compatible with it. With the given
256: symmetry, we choose the metric on the orbits of the ${\rm SO(3)}$
257: action to be the unit 2-sphere metric,
258: $\md\vt^2+\sin^2\vt\md\vp^2$ in polar coordinates $\vt,\vp$.
259: Locally, the corresponding co-triad elements $\md\vt$, and
260: $\sin\vt\md\vp$ can be completed to an orthonormal co-triad by a
261: third form $\alpha$. Thanks to the Maurer--Cartan relations for
262: the symmetry group, $\alpha$ must be closed and therefore exact on
263: the given topology, whence it defines a third adapted coordinate
264: $x$ via $\alpha=\md x$. In these coordinates the background metric
265: has line element
266: %
267: $$\md s_o^2= \md x^2+\md\vt^2+\sin^2\vt\md\vp^2$$
268: %
269: and determinant $\q=\sin^2\vt$.
270:
271: In connection dynamics, the canonically conjugate pair consists of
272: fields $(A_a^i, E^a_i)$ of an ${\rm SU(2)}$ connection 1-form
273: $A_a^i$ and a (possibly degenerate) triad $E^a_i$ of density
274: weight 1 on $M$. This pair is invariant under the symmetry group
275: if it satisfies:
276: %
277: \be \mathcal{L}_\xi\, A = D\Lambda \quad {\rm and} \quad
278: \mathcal{L}_\xi E = [E, \Lambda] \label{sym} \ee
279: %
280: for any symmetry vector field $\xi^a$ and some generator
281: $\Lambda^i$ of local, ${\rm SU(2)}$ gauge transformations (which
282: may depend on $\xi^a$). One can verify that any such symmetric
283: pair is gauge equivalent to fields of the form:%
284: %
285: \footnote{More precisely, invariant connections \cite{kn,b}
286: carry a non-negative, integer-valued topological charge which we
287: have set equal to one in the above expressions. (In general, it
288: would multiply the last contribution to (\ref{sym1}); see, e.g.,
289: \cite{bk}.) Each value of this charge gives rise to an independent
290: sector of invariant connections. However, only the sector used
291: here allows non-degenerate triads invariant under (\ref{sym}) with
292: {\em the same} local gauge transformation in $A$ and $E$. That $A$
293: in (\ref{sym1}) satisfies (\ref{sym}) can be verified as follows.
294: If we denote the three generators of the ${\rm SO(3)}$ action on
295: $M$ by
296: $X=\sin\vp\partial_{\vt}+\cot\vt\cos\vp\partial_{\vp}$,
297: $Y=-\cos\vp\partial_{\vt}+\cot\vt\sin\vp\partial_{\vp}$, and
298: $Z=\partial_{\vp}$, and the generator of the $\R$-action by $t^a$,
299: then $\mathcal{L}_X A = D(\cos\vp/\sin\vt)\tau_3$,\,
300: $\mathcal{L}_Y A = D (\sin\vp/\sin\vt)\tau_3$,\, $\mathcal{L}_Z A
301: = 0$ and $\mathcal{L}_t A = 0$. Invariance of $E$ in (\ref{sym2})
302: follows analogously, or by noting that $E^a_i\delta A_a^i$ defines
303: a gauge invariant and (spatially) constant 1-form on phase space.}
304: %
305: \ba A &=& \tilde{c}\tau_3\md
306: x+(\tilde{a}\tau_1+\tilde{b}\tau_2)\md\vt
307: +(-\tilde{b}\tau_1+\tilde{a}\tau_2)\sin\vt\md\vp+
308: \tau_3 \cos\vt \md\vp \label{sym1}\\
309: E &=& \tilde{p}_c\tau_3\sin\vt\frac{\partial}{\partial
310: x}+(\tilde{p}_a\tau_1+\tilde{p}_b\tau_2)
311: \sin\vt\frac{\partial}{\partial\vt}
312: +(-\tilde{p}_b\tau_1+\tilde{p}_a\tau_2)\frac{\partial}{\partial\vp}\,.
313: \label{sym2} \ea
314: %
315: where $\tilde{a},\tilde{b},\tilde{c}$ are real constants and
316: $\tau^j$ is the standard basis in ${\rm su(2)}$, satisfying
317: $[\tau_i,\, \tau_j] = \epsilon_{ij}{}^k \tau_k$. However, this
318: does not exhaust the freedom to perform local ${\rm SU(2)}$ gauge
319: transformations entirely: There is still the freedom to perform a
320: \emph{global} $\SU(2)$ transformation along $\tau_3$ which rotates
321: the `vectors' $(\tilde{a}, \tilde{b})$ and $(\tilde{p_a},
322: \tilde{p_b})$. We will return to this freedom in the next
323: sub-section.
324:
325: From the invariant triad density (\ref{sym2}) we can derive the
326: corresponding co-triad $\omega$:
327: %
328: \be\label{cotriad} \omega = \omega_c \tau_3 \md x +
329: (\omega_a\tau_1 + \omega_b\tau_2) \md \vt + (-\omega_b \tau_1 +
330: \omega_a \tau_2) \sin\vt \md\vp\ee
331: %
332: where
333: %
334: \be \omega_c = \frac{{\rm sgn}\tilde{p}_c\,
335: \sqrt{\tilde{p}^2_a+\tilde{p}^2_b}}{\sqrt{|\tilde{p}_c|}};\quad
336: \omega_b = \frac{\sqrt{|\tilde{p}_c|}\, \tilde{p}_b}
337: {\sqrt{\tilde{p}^2_a+\tilde{p}^2_b}}; \quad {\rm and} \,\,
338: \omega_a = \frac{\sqrt{|\tilde{p}_c|}\,\tilde{p}_a}
339: {\sqrt{\tilde{p}^2_a+\tilde{p}^2_b}}\, . \ee
340: %
341: The co-triad in turn determines its spin-connection, the ${\rm
342: su(2)}$-valued 1-form $\Gamma$. As one might expect from the
343: structure of the triad, $\Gamma$ turns out to be the standard
344: `magnetic monopole' connection:
345: %
346: \begin{equation} \label{Gamma}
347: \Gamma= \cos\vt\tau_3\, \md\vp\,.
348: \end{equation}
349: %
350: Consequently, the ${\rm su(2)}$-valued extrinsic curvature 1-form
351: $K$ is given by:
352: %
353: \begin{equation} \label{K}
354: \gamma K : =A - \Gamma= \tilde{c}\tau_3\md x +
355: (\tilde{a}\tau_1+\tilde{b}\tau_2)\md\vt
356: + (-\tilde{b}\tau_1+\tilde{a}\tau_2)\sin\vt\md\vp\,.
357: \end{equation}
358: %
359: Finally the curvature $F_{ab}$ of $A_a$ can be easily computed.
360: Its dual, the ${\rm su(2)}$-valued vector density $B^a\, :=\,
361: \frac{1}{2} \eta^{abc}F_{bc}$ has the same invariant form as $E^a$:
362: %
363: \begin{equation} \label{B}
364: B=(\tilde{a}^2+\tilde{b}^2-1)\tau_3\sin\vt\frac{\partial}{\partial
365: x}+(\tilde{a}\tau_1+\tilde{b}\tau_2)\tilde{c}
366: \sin\vt\frac{\partial}{\partial\vt}
367: +(-\tilde{b}\tau_1+\tilde{a}\tau_2)\tilde{c}\frac{\partial}{\partial\vp}
368: \end{equation}
369: %
370: These expressions will be used in sections \ref{s3} and \ref{s4}.
371:
372:
373: \subsection{The reduced phase space}
374: \label{s2.2}
375:
376: The symmetry reduction procedure led us to a 6-dimensional reduced
377: phase space $\tilde{\bf \Gamma}$ with coordinates $(\tilde{a},
378: \tilde{b}, \tilde{c};\, \tilde{p}_a, \tilde{p}_b, \tilde{p}_c)$.
379: Let us begin by computing the symplectic structure $\tilde{\bf
380: \Omega}$ on it. The basic idea, as in all symmetry reductions, is
381: to pull-back the symplectic structure of the full theory to the
382: symmetry reduced phase space. However, since the full symplectic
383: structure involves an integral over $M$ and since the fields of
384: interest are homogeneous also in the non-compact
385: $\mathbb{R}$-direction, as usual, we are led to consider only a
386: finite interval $\cal{I}$ in the $\mathbb{R}$ direction. Let the
387: length of this interval (w.r.t. the fiducial metric $\q_{ab}$) be
388: $L_o$. Then, the symplectic structure on the reduced phase space
389: is given by:
390: %
391: \begin{equation}
392: \tilde{\bf \Omega} = \frac{L_o}{2\gamma G}(2\md\tilde{a}\wedge\md
393: \tilde{p}_a+ 2\md\tilde{b}\wedge\md \tilde{p}_b+ \md
394: \tilde{c}\wedge\md \tilde{p}_c)
395: \end{equation}
396: %
397: where $\gamma$ is the Barbero--Immirzi parameter and $G$ the
398: gravitational constant. (The volume of the $\S^2$ orbits of the
399: ${\rm SO(3)}$ action, defined by the fiducial metric, is the standard
400: one, $4\pi$.)
401:
402: Because of the form (\ref{sym1}) and (\ref{sym2}) of invariant
403: connections and triads, the vector constraint is automatically
404: satisfied. However, because of the residual, global ${\rm SU(2)}$
405: gauge freedom mentioned in section \ref{s2.1}, the Gauss
406: constraint is not. Inserting the invariant connection and triad
407: into the Gauss constraint we obtain
408: %
409: \begin{equation}
410: \Gauss=\tilde{a}\tilde{p}_b-\tilde{b}\tilde{p}_a=0
411: \end{equation}
412: %
413: which generates simultaneous rotations of the pairs
414: $(\tilde{a},\tilde{b})$ and $(\tilde{p}_a,\tilde{p}_b)$. Thus,
415: only the `norms' $\sqrt{\tilde{a}^2+\tilde{b}^2}$,
416: $\sqrt{\tilde{p}_a^2+\tilde{p}_b^2}$ and the `scalar product'
417: $(\tilde{a}\tilde{p}_a+\tilde{b}\tilde{p}_b)$ are gauge invariant.
418: We will fix this gauge freedom as follows. If
419: $(\tilde{a},\tilde{b})=0$, we rotate the triad components such
420: that $\tilde{p}_a=0$. Otherwise we rotate $(\tilde{a},\tilde{b})$
421: such that $\tilde{a}=0$ which implies $\tilde{b}\not=0$. Then,
422: $\Gauss=0$ implies $\tilde{p}_a=0$. There is still a residual
423: gauge freedom\,\, $\Pi_b\colon (b, p_b) \rightarrow (-b, -p_b)$\,\,
424: which changes the signs of $b$ and $p_b$ simultaneously. This is
425: just the parity transformation in the $p_b$ variable. One can
426: either retain this freedom and ensure that all the final
427: constructions are invariant with respect to this `$b$-reflection'
428: \emph{or} eliminate it by a gauge choice such as $\tilde{p}_b \ge
429: 0$. This gauge choice turns out to be particularly convenient for
430: classical dynamics. In the quantum theory, on the other hand, it
431: is more natural to retain the freedom at first and then ask that
432: physical states be invariant under the parity operator
433: $\hat{\Pi}_b$ implementing this transformation. Therefore, we will
434: allow both possibilities. In either case the 4-dimensional phase
435: space carries a global chart $(\tilde{b},\tilde{c};
436: \tilde{p}_b,\tilde{p}_c)$ (where $\tilde{p}_b \ge0$ if the
437: `$b$-parity gauge' is fixed.) These coordinates are subject to the
438: scalar or the Hamiltonian constraint, discussed below.
439:
440: We first note that these variables and, because of its explicit
441: dependence on $L_o$, the symplectic structure $\tilde{\bf \Omega}$
442: depend on the fiducial metric. It is convenient to remove this
443: dependence by rescaling the variables in a manner that is
444: motivated by their scaling properties,
445: %
446: \begin{equation}
447: (b,c):=(\tilde{b},L_o\tilde{c})\quad,\quad
448: (p_b,p_c):=(L_o\tilde{p}_b,\tilde{p}_c)\,.
449: \end{equation}
450: %
451: We now have a one-to-one parametrization of the gauge fixed phase
452: space ${\bf \Gamma} $ by $(b,p_b,c,p_c)$ with symplectic structure
453: %
454: \begin{equation} \label{symp}
455: {\bf \Omega} = \frac{1}{2\gamma G}(2\md b\wedge\md p_b+\md
456: c\wedge\md p_c)\,.
457: \end{equation}
458: %
459: For later purposes let us express the volume of our elementary
460: cell ${\cal I}\times \S^2$ and areas bounded by our preferred
461: family of curves as functions on the reduced phase space. The
462: volume is given by:
463: %
464: \begin{equation} \label{vol}
465: V=\int\md^3x\sqrt{\left|\det E\right|}=4\pi L_o
466: \sqrt{|\tilde{p}_c|} |\tilde{p}_b| = 4\pi\sqrt{|p_c|} |p_b|\,.
467: \end{equation}
468: %
469: The three surfaces $S_{x,\vp}$, $S_{x,\vt}$ and $S_{\vt,\vp}$ of
470: interest are respectively bounded by the interval ${\cal I}$ and
471: the equator, ${\cal I}$ and a great circle along a longitude, and
472: the equator and a longitude (so that $S_{\vt,\vp}$ forms a quarter
473: of the sphere $\S^2$). Their areas are given by:
474: %
475: \be \label {area}
476: A_{x, \vt} = 2\pi |p_b|, \quad A_{x, \vp} = 2 \pi |p_b|,
477: \quad{\rm and} \quad A_{\vt,\vp} = \pi |p_c|\,. \ee
478:
479: Finally, let us consider the Hamiltonian constraint and classical
480: dynamics. On the reduced phase space ${\bf \Gamma}$, the
481: constraint functional of the full theory
482: %
483: \begin{equation} \label{fullham}
484: \ham= \int\md x^3 N e^{-1}\, [\epsilon_{ijk} F_{ab}^i E^a_jE^b_k-
485: 2(1+\gamma^2) K_{[a}^iK_{b]}^j E_i^aE_j^b]\, ,
486: \end{equation}
487: %
488: where $e:= \sqrt{\left|\det E\right|}\,{\rm sgn}(\det E)$, reduces
489: to
490: %
491: \begin{eqnarray} \label{ksham}
492: \ham = -\frac{8\pi N}{\gamma^2}\frac{\sgn p_c}{\sqrt{|p_c|}p_b}
493: \left[(b^2+\gamma^2)p_b^2+
494: 2cp_cbp_b\right]\, ,
495: \end{eqnarray}
496: %
497: where, as is usual in the homogeneous models, we have chosen a
498: constant lapse function $N$. To simplify the equations of motion,
499: it is convenient%
500: %
501: \footnote{With this choice, the points $b=0$ or $p_c =0$ have to
502: be excised in the discussion of dynamics. The explicit form of the
503: solutions below shows that this is not overly restrictive since on
504: classical solutions these points correspond to the singularity or
505: the horizon.}
506: %
507: to choose $N = \frac{\gamma {\rm sgn}p_c\, \sqrt{|p_c|}}{16\pi G
508: b}$. Then the Hamiltonian constraint becomes $\ham:=-(2\gamma
509: G)^{-1}[(b^2+\gamma^2)p_b/b+2cp_c]$, yielding the following
510: equations of motion:
511: %
512: \begin{eqnarray*}
513: \dot{b} = -\frac{1}{2}(b+\gamma^2b^{-1}), &\quad\quad&
514: \dot{p}_b = \frac{1}{2}(p_b-\gamma^2b^{-2}p_b)\\
515: \dot{c} = -2c, &\quad\quad&
516: \dot{p}_c = 2p_c\, ,
517: \end{eqnarray*}
518: %
519: where the `dot' denotes derivative with respect to the affine
520: parameter of the Hamiltonian vector field. The equations for
521: $(c,p_c)$ can easily be solved by $c(\T)=c_0e^{-2\T}$ and
522: $p_c(\T)=p_c^{(0)}e^{2\T}$. For the other components we obtain
523: $b(\T)=\pm\gamma\sqrt{e^{-(\T-\T_0)}-1}$ for $b$ and, using this
524: result, $p_b(T)=p_b^{(0)}\sqrt{e^{\T+\T_0} -e^{2\T}}$. If we
525: introduce the new time parameter $t:=e^{\T}$ and the constant
526: $m:=\frac{1}{2}e^{\T_0}$, this gives
527: %
528: \begin{eqnarray}
529: b(t) = \pm\gamma\sqrt{(2m-t)/t}, &\quad\quad&
530: p_b(t) = p_b^{(0)}\sqrt{t(2m-t)}\\
531: c(t) = \mp \gamma m p_b^{o}\, t^{-2}\label{3}, &\quad\quad&
532: p_c(t) = \pm \,t^2\label{4}
533: \end{eqnarray}
534: %
535: where we have fixed the multiplicative constant in the expressions of
536: $p_c(t)$ and $c(t)$, respectively, by requiring that $|p_c|$ be the
537: geometric radius of the 2-sphere orbits of the ${\rm SO(3)}$ symmetry
538: and by using the Hamiltonian constraint. Projections to the $p_b-p_c$
539: plane of typical trajectories are shown in figure \ref{traj2}. It is
540: clear from the figure that $p_c$ can be taken to be the `internal time
541: parameter'. In section \ref{s4.1} this interpretation will let us
542: regard the quantum Hamiltonian constraint as an `evolution equation'
543: with respect to an internal time parameter defined by the eigenvalues
544: of the operator $\hat{p}_c$.
545:
546: \begin{figure}
547: \begin{center}
548: \includegraphics[width=3in]{traj2.eps}
549: \caption{\footnotesize Dynamical trajectories in the $p_b-p_c$
550: plane. Each trajectory reaches the maximum value $m$ of $p_b$ and
551: meets the $p_c =0$ axis only at the point where $p_b$ also
552: vanishes (i.e., at the `origin'). Solutions related by the
553: reflection-symmetry $p_c \rightarrow -p_c$ define the same metric
554: but carry triads of opposite orientation. For simplicity the
555: `$b$-parity gauge' is fixed by requiring $p_b \ge 0$.}
556: \label{traj2}
557: \end{center}
558: \end{figure}
559:
560: Note that we have a 2-parameter family of solutions, labeled by
561: $m$ and $p^{(0)}_b$, as one would expect from the fact that the
562: reduced phase space is 2-dimensional. However, in a space-time
563: description we expect only a 1-parameter family, labeled by $m$.
564: This discrepancy can be traced back to the standard tension
565: between the Hamiltonian and space-time notions of gauge (for a
566: detailed discussion in the context of Bianchi models, see
567: \cite{as}, and in spherically symmetric models, \cite{kt,k}). In
568: the Hamiltonian description, rescalings of $p^{(0)}_b$ do not
569: correspond to gauge because they are not generated by any
570: constraint. In the space-time description, on the other hand, they
571: can be absorbed by a rescaling of the coordinate $x$. Therefore,
572: to make contact with the space-time description, let us fix
573: $p^{(0)}_b =1$. Then, the above solution to the evolution
574: equations defines a 1-parameter family of 3-metrics:
575: %
576: \be q_{ab}(t) = \left(\frac{2m}{t} -1\right) \nabla_a x \nabla_b x
577: + t^2 (\nabla_a\vt \nabla_b \vt + \sin^2\vt \nabla_a
578: \vp\nabla_b\vp ) \ee
579: %
580: defining precisely the Schwarzschild solution of mass $m$,
581: interior to the horizon. It will be useful to note that, in this
582: solution $p_c>0$, $p_b=0$ at the horizon ($t=2m$) and $p_c =0$ and
583: $b,c$ diverge at the singularity.
584:
585: We will see in the next section that it is convenient to regard
586: the space $\M$ spanned by $p_b$ and $p_c$ as the mini-superspace
587: on which quantum wave functions are defined. (In that discussion,
588: we will not fix the `$b$-parity gauge'; so $p_c$ \emph{and} $p_b$
589: take values on the entire real line.) Let us therefore interpret
590: various regions in $\M$. The triad is non-degenerate everywhere
591: except when $p_b=0$ or $p_c =0$. However, the two degeneracies are
592: of very different geometric and physical origin. Geometrically
593: (i.e., independent of classical dynamics) $p_c=0$ separates two
594: regions with triads of opposite orientations, given by $\sgn\det
595: E=\sgn(p_c p_b^2)$. On the other hand, as (\ref{cotriad})
596: (together with our gauge choice which ensures $a=p_a =0$) shows,
597: the orientation would not have changed even if we had extended
598: $\M$ across $p_b=0$, allowing for negative values for $p_b$. More
599: importantly, while the co-triad (\ref{cotriad}) remains smooth and
600: only becomes degenerate along $p_b=0$, it diverges along $p_c =0$,
601: signaling the classical singularity. This suggests that the line
602: $p_b =0$ corresponds to the horizon and $p_c =0$ to the
603: singularity. This is confirmed from physical considerations
604: involving classical dynamics. Classical solutions can be smoothly
605: extended to the line $p_b=0$ which is reached for $t=2m$. The
606: curvature $B$ remains well-defined there (see (\ref{B})). The
607: other line, $p_c=0$, is very different. Along classical solutions,
608: the line $p_c=0$ is approached as $t\to 0$, whence the components
609: (and invariants) of curvature $B$ also diverge. This line
610: corresponds to the singularity and cannot be crossed by any
611: classical dynamical trajectory. Finally, note that since the line
612: $p_c=0$ separates the two regions of $\M$, where only the triad
613: orientation is reversed, the two get identified in
614: geometrodynamics. Therefore the singularity would lie at a
615: boundary of the mini-superspace in geometrodynamics, making the
616: geometrical meaning of `evolution across the singularity' obscure.
617: Furthermore, since the Arnowitt--Deser--Misner (ADM) variables are
618: based on covariant metrics (rather than contravariant triad
619: densities), as in other homogeneous models \cite{mb6}, one of the
620: ADM variables becomes {\em infinite} at the classical singularity.
621: This further complicates a discussion of the singularity structure
622: and its resolution.
623:
624:
625: \section{Quantum Kinematics}
626: \label{s3}
627:
628: This section is divided into two parts. In the first we introduce
629: the basic quantization procedure and in the second we discuss
630: quantum geometry with emphasis on the classical singularity.
631:
632: \subsection{Basics}
633: \label{s3.1}
634:
635: As in the isotropic model \cite{abl}, we will follow the general
636: procedure used in the full theory \cite{alrev,crbook,ttrev}. Thus
637: the elementary configuration variables will be given by holonomies
638: along curves in $M$ and the momenta by fluxes of triads along
639: 2-surfaces in $M$. However, because of symmetry reduction, one
640: need not consider all (piecewise analytic) curves or surfaces; a
641: judiciously chosen, smaller subset suffices to obtain a set of
642: functions that is sufficiently large to separate points of the
643: reduced phase space.
644:
645: Let us begin with the holonomies. We will restrict ourselves to
646: three sets of curves: those along the $\R$ direction of $M$ with
647: oriented length $\nu L_o$; those along the equator of $\S^2$ with
648: oriented length $\mu$; and those along the longitudes of $\S^2$
649: also with oriented length $\mu$, where all lengths and
650: orientations are defined using the fiducial triad. (Thus $\nu,
651: \mu$ are positive if the tangent to the respective curves are
652: parallel to the triad vectors and negative if they are
653: anti-parallel.) Holonomies along these curves suffice to
654: completely determine any invariant connection (\ref{sym1}).%
655: %
656: \footnote{In the fully gauge fixed setting, we could omit
657: $h_{\vt}$ since it is related to $h_{\vp}$ by conjugation with
658: $\exp\,\tau_3$. However, we will use a more `democratic' approach
659: which is applicable also when $a$ is not made to vanish by gauge
660: fixing.}
661: %
662: \begin{eqnarray}
663: h^{(\nu)}_x(A) &=& \exp\int_0^{\nu L_o}\md x \tilde{c}\tau_3=
664: \cos\frac{\nu c}{2}+ 2\tau_3\sin\frac{\nu c}{2} \label{hol1}\\
665: h^{(\mu)}_{\vp}(A) &=& \exp-\int_0^{\mu}\md\vp \tilde{b}\tau_1=
666: \cos\frac{\mu b}{2}- 2\tau_1\sin\frac{\mu b}{2} \label{hol2}\\
667: h^{(\mu)}_{\vt}(A) &=& \exp\int_0^{\mu}\md\vt \tilde{b}\tau_2=
668: \cos\frac{\mu b}{2}+ 2\tau_2\sin\frac{\mu b}{2}\, .\label{hol3}
669: \end{eqnarray}
670: %
671: Matrix elements of these holonomies are functions of the reduced
672: connection and constitute our configuration variables. Elements of
673: the algebra they generate are almost periodic functions of $b$ and
674: $c$ of the form $f(b,c) = \sum_{\mu,\nu}\, f_{\mu\nu} \exp
675: {\frac{i}{2}\, (\mu b + \nu c)}$, where $f_{\mu,\nu} \in
676: {\mathbb{C}}$, $\mu,\nu \in \mathbb{R}$ and the sum extends over a
677: finite set. This algebra is the Kantowski-Sachs analog of the
678: algebra of cylindrical functions in the full theory
679: \cite{alrev,crbook,ttrev}. We will therefore denote it by $\cyl$.
680: Consider the $C^\star$ algebra obtained by completing $\cyl$ using
681: the sup norm. The quantum configuration space $\Ab$ is the
682: Gel'fand spectrum of this algebra. From the structure of the
683: algebra, it follows that the spectrum is naturally isomorphic to
684: the Bohr compactification $\bar{\R}_{\rm Bohr}^2$ of the Abelian
685: group $\R^2$ \cite{afw,jw}. (Recall that in the isotropic case
686: \cite{abl}, the quantum configuration space is isomorphic to
687: $\bar{\R}_{\rm Bohr}$.) The Hilbert space $\tilde\H$ is obtained
688: by the Cauchy completion of $\cyl$ with respect to the natural
689: Haar measure $\mu_o$ on the Abelian group $\bar{\R}^2_{\rm Bohr}$;
690: $\tilde\H = L^2(\bar{\R}_{\rm Bohr}^2,\md\mu_o)$. Employing the
691: standard bra-ket notation, we can define a basis $|\mu,\nu\rangle$
692: in $\tilde\H$ via:
693: %
694: \begin{equation} \label{basis}
695: \langle b,c|\mu,\nu\rangle = e^{\frac{i}{2}\,(\mu b+\nu c)} \qquad
696: \mu,\nu\in{\R}\, .
697: \end{equation}
698: %
699: This is an orthonormal basis:
700: %
701: \be \langle \mu^\prime, \nu^\prime|\mu,\nu\rangle =
702: \delta_{\mu^\prime\, \mu}\, \delta_{\nu^\prime\, \nu} \, \ee
703: %
704: where, on the right side, we have the Kronecker symbol, rather
705: than the Dirac delta distribution. Thus, each basis vector is
706: normalizable and has unit norm.
707:
708: As one would expect, the configuration ---i.e., holonomy---
709: operators $\hat{h}_x^{(\nu)},\, \hat{h}_\vp^{(\mu)},\,
710: \hat{h}_\vt^{(\mu)}$ operate on $\tilde\H$ by multiplication.
711: However, as in the full theory, these operators are not required
712: to be weakly continuous in parameters $\mu$ and $\nu$, whence
713: there are no operators $\hat{b}, \hat{c}$ corresponding to the
714: connection itself. consequently, although we have only a finite
715: number of degrees of freedom, the von-Neumann uniqueness theorem
716: is inapplicable and this quantum theory is inequivalent to a
717: standard `Schr\"odinger quantization' (for a further discussion,
718: see \cite{afw,jw}). We will see that, as in the isotropic model
719: \cite{abl}, this inequivalence has important consequences.
720:
721: For the momentum operators we consider fluxes of triads along
722: preferred 2-surfaces. Apart from fixed kinematical factors, they
723: are given by components $p_b,p_c$ of triads which, in view of the
724: symplectic structure (\ref{symp}), are represented by operators%
725: %
726: \footnote{In this paper, we use the standard quantum gravity
727: convention, $\lp^2 = G\hbar$. Unfortunately, this is different
728: from the convention $\lp^2 = 8\pi G\hbar$ used in most of the loop
729: quantum cosmology literature. Hence care should be exercised while
730: comparing detailed numerical factors.}
731: %
732: \be \hat{p}_b = -i{\gamma\lp^2}\,\frac{\partial}{\partial b},
733: \quad\quad \hat{p}_c = -2i\gamma\lp^2 \frac{\partial}{\partial c}\,.
734: \ee
735: %
736: %\begin{eqnarray} \label{triadop}
737: % \hat{p}_b &=& -i\frac{\gamma\lp^2}{8\pi}\frac{\partial}{\partial
738: %b}\\
739: % \hat{p}_c &=& -i\frac{\gamma\lp^2}{4\pi}\frac{\partial}{\partial
740: %c}
741: %\end{eqnarray}
742: %
743: Their eigenstates are the basis states (\ref{basis}),
744: %
745: \be \hat{p}_b|\mu,\nu\rangle =
746: \textstyle{\frac{1}{2}}\,\gamma\lp^2\, \mu |\mu,\nu\rangle,\qquad
747: \hat{p}_c|\mu,\nu\rangle = \gamma\lp^2\, \nu
748: |\mu,\nu\rangle\, . \ee
749:
750: However, we still have to incorporate the residual gauge freedom
751: which corresponds to a parity reflection in the $b$ degree of
752: freedom. Therefore, only those states in $\tilde\H$ which are
753: invariant under the parity operator $\hat\Pi_b\colon |\mu,\nu\rangle
754: \rightarrow |-\mu,\nu\rangle$ can belong to the kinematical
755: Hilbert space $\H$. A basis in $\H$ is thus given by:
756: %
757: \be \frac{1}{\sqrt{2}}\, [|\mu, \nu\rangle + |-\mu, \nu\rangle
758: ]\ee
759: %
760: Finally, we express the volume operator in terms of triad
761: operators $\hat{p}_b, \hat{p}_c$. Recall that the region of $M$
762: under consideration is of the type ${\cal I} \times \S^2$, where
763: ${\cal I}$ has length $L_o$ with respect to the fiducial metric.
764: From the classical expression (\ref{vol}) of the volume $V$ of
765: this region, it follows that the operator $\hat{V}$ is given by
766: $\hat{V}=4\pi |\hat{p}_b|\, \sqrt{|\hat{p}_c|}$. It is diagonal in
767: our $|\mu, \nu\rangle$ basis and the eigenvalues are:
768: %
769: \begin{equation}
770: V_{\mu\nu}=2\pi\, \gamma^{3/2}\, |\mu|\, \sqrt{|\nu|}\,\, \lp^3\, .
771: \end{equation}
772: %
773: As in the isotropic case, the volume of our cell ${\cal I} \times
774: \S^2$ with respect to the fiducial metric can be large, but its
775: volume in the `elementary' state $\exp \frac{i}{2}\,(b+c)$, is of
776: Planck size.
777:
778:
779: \subsection{Quantum geometry}
780: \label{s3.2}
781:
782: As in the isotropic case the triad operators $\hat{p}_b$,
783: $\hat{p}_c$ commute whence, in contrast to the full theory, one can
784: construct the triad representation. In this sub-section we will
785: study the quantum Riemannian geometry using this representation.
786: In the next sub-section we will see that the representation is
787: also convenient for dynamics, for it suggests an intuitive
788: interpretation for the action of the Hamiltonian constraint.
789:
790: Let us then expand a general state in terms of eigenstates of the
791: triad operators, $|\psi\rangle= \sum_{\mu\nu}\psi_{\mu\nu}
792: |\mu,\nu\rangle$, and use the coefficients $\psi_{\mu\nu}$ to
793: represent the state. This wave function $\psi_{\mu\nu}$ is
794: supported on the mini-superspace $\mathcal{M}$ coordinatized by
795: $p_b$ and $p_c$. In the classical theory, already at the kinematical
796: level, the lines $p_b =0$ and $p_c =0$ are special: the co-triad
797: (\ref{cotriad}) becomes degenerate along $p_b=0$ (and the line
798: represents the horizon) while it diverges on the line $p_c=0$ (and
799: this line represents the singularity). It is then natural to
800: examine wave functions $\psi_{\mu\nu}$ which are supported on
801: lines $\mu=0$ and $\nu=0$ and ask, already at the kinematical
802: level, for the nature of quantum geometry they represent.
803:
804: Let us consider the co-triad component $\omega_c \equiv
805: \sgn(p_c)\, |p_b|/\sqrt{|p_c|}$ along $\md x$ (see
806: (\ref{cotriad})). Since it involves an inverse power of $p_c$ and
807: since there exist normalizable kets $|\mu, \nu=0\rangle$ of the
808: operator $\hat{p}_c$ with zero eigenvalue, the naive operator
809: obtained by replacing $p_b$ and $p_c$ by corresponding operators
810: is not even densely defined, let alone self-adjoint. To define
811: $\hat\omega_c$, therefore, a new strategy is needed. We will
812: follow what is by now a standard procedure, adapted from
813: Thiemann's analysis in the full theory \cite{tt}. The first step
814: is to express $\omega_c$ in terms of the elementary variables
815: which do have unambiguous quantum analogs ---holonomies and
816: positive powers of $p_b, p_c$--- and Poisson brackets between
817: them. On the classical phase space, we have the exact equality:
818: %
819: \begin{equation} \label{omegaccomm}
820: \omega_c=\frac{1}{2\pi\gamma
821: G}\tr\left(\tau_3 h_x \{h_x^{-1},\,V\}\right)
822: \end{equation}
823: %
824: %
825: %\begin{equation}
826: % e_c=\frac{1}{2\pi\gamma
827: %G\L}\tr\left(\tau_3h_x^{(\L)}\{h_x^{(\L)-1},V\}\right)
828: %\end{equation}
829: %
830: where $h_x$ is the holonomy along the interval ${\cal I}$, i.e.,
831: along the edge of length $L_o$ (with respect to the fiducial
832: metric) we fixed in our construction of the phase space. Then,
833: replacing $h_x$ and $V$ by their unambiguous quantum analogs and
834: the Poisson bracket by $1/i\hbar$ times the commutator, we obtain
835: %
836: \begin{eqnarray}
837: \hat{\omega}_c &=& -\frac{i}{2\pi \gamma\lp^2}\, \tr\left(\tau_3
838: \hat{h}_x[\hat{h}_x^{-1},\, \hat{V}]\right)\nonumber\\
839: &=& -\frac{i}{2\pi \gamma\lp^2}\left(\sin\frac{
840: c}{2}\hat{V}\cos\frac{c}{2}- \cos\frac{c}{2}\hat{V}\sin\frac{
841: c}{2}\right)\,.
842: \end{eqnarray}
843: %
844: %\begin{eqnarray}
845: % \hat{e}_c &=& -\frac{4i}{\gamma\L\lp^2} \tr\left(\tau_3
846: %\hat{h}_x^{(\L)}[\hat{h}_x^{(\L)-1},\hat{V}]\right)\nonumber\\
847: % &=& -\frac{4i}{\gamma\L\lp^2}\left(\sin\frac{\L
848: %c}{2}\hat{V}\cos\frac{\L c}{2}- \cos\frac{\L c}{2}\hat{V}\sin\frac{\L
849: %c}{2}\right)\,.
850: %\end{eqnarray}
851: %
852: It turns out that this operator is diagonalized by our basis
853: $|\mu, \nu\rangle$ of (\ref{basis}),
854: %
855: \begin{equation}
856: \hat{\omega}_c |\mu,\nu\rangle = \frac{1}{4\pi \gamma\lp^2}
857: (V_{\mu,\nu+ 1}-V_{\mu,\nu- 1})\, |\mu,\nu\rangle =
858: \frac{\sqrt{\gamma}}{2}\lp\, |\mu|\, (\sqrt{|\nu+ 1|}-\sqrt{|\nu-
859: 1|})\,\, |\mu,\nu\rangle\,.
860: \end{equation}
861:
862:
863: For large $\nu$, far away from the classical singularity, these
864: eigenvalues are very close to the classical expectation
865: $|\omega_c|= |p_b|/\sqrt{|p_c|}$. This is even true for small
866: $\mu$, i.e.\ we do not need to be far from the horizon. Closer to
867: the classical singularity $p_c=0$, on the other hand, the behavior
868: becomes significantly different from the classical one. At the
869: classical singularity itself, where $|p_b|/\sqrt{|p_c|}$ diverges,
870: the eigenvalue of $\hat{\omega}_c$ is zero.%
871: %
872: \footnote{Exact classical identities such as (\ref{omegaccomm})
873: for objects containing inverse powers are available only in
874: homogeneous models. More generally, the bracket between holonomies
875: and volume is more complicated \cite{mb10}. These complications
876: lead to additional correction terms and the resulting operator is
877: no longer bounded on eigenspaces of $\hat{p}_b$.}
878: %
879:
880: \emph{Remark:} In the above definition of $\hat{\omega}_c$ we
881: needed a holonomy in the $x$-direction. We chose to evaluate it
882: using the interval ${\cal I}$ of length $L_o$ with respect to the
883: fiducial metric. While this choice is natural because this
884: interval appears already in the construction of the phase space,
885: we could have used an interval of length $L_o\delta$ for some
886: $\delta$. Then the resulting operator would also have been bounded
887: on the entire Hilbert space $\tilde\H$,\, and $|\mu, \nu
888: =0\rangle$ would again have been eigenvectors of the operator with
889: zero eigenvalue. Thus, the qualitative properties of the operator
890: are insensitive to $\delta$. Furthermore, from general
891: considerations one can argue \cite{abl} that $\delta$ should be of
892: the order of $1$. However, the value of the upper bound of the
893: spectrum does depend on $\delta$. Therefore, as in the isotropic
894: model, the precise numerical coefficient in this bound should not
895: be attributed physical significance.
896:
897: Let us now turn to the second triad component $\omega_b = {\rm
898: sgn} p_b\,\, \sqrt{|p_c|}$. Since $\sqrt{|\nu|}$ is a well-defined
899: function of $\nu$ on the entire spectrum of $\hat{p}_c$, we can
900: quantize $\omega_b$ directly. This operator is again diagonal in
901: our basis $|\mu, \nu\rangle$ and its eigenvalues are given by
902: %
903: \begin{equation}
904: \hat{\omega}_b|\mu,\nu\rangle=
905: \sqrt{\gamma}\lp\sgn(\mu)\sqrt{|\nu|}\,|\mu,\nu\rangle\,.
906: \end{equation}
907: %
908: Since the operator is quantized directly, the eigenvalues are just
909: the ones one would expect from the classical expression
910: (\ref{cotriad}) of $\omega_b$.
911:
912: The properties of co-triad operators suggest that the lines
913: $\mu=0$ and $\nu=0$ have very different features also in quantum
914: geometry. No quantum effects are manifest at $\mu=0$ which
915: corresponds to the classical horizon. Near and on the line $\nu=0$
916: which represents the singularity, on the other hand, quantum
917: effects are large. In particular, they remove the classical
918: singular behavior of the co-triad. We emphasize, however, that
919: even in homogeneous models, the ultimate test as to whether or not
920: a singularity persists upon quantization can only come from
921: studying the dynamics. The key questions are: Is quantum dynamics
922: well-defined and deterministic across the classical singularity?
923: And, does this come about without significant quantum corrections
924: to geometry near the horizon? Only quantum dynamics will tell us
925: if the indication provided by the properties of the co-triad
926: operators are borne out.
927:
928: \section{Quantum Dynamics}
929: \label{s4}
930:
931: This section is divided into three parts. In the first we present
932: the strategy, in the second we obtain the quantum Hamiltonian
933: constraint, and in the third we use it to discuss the consequent
934: resolution of the classical singularity.
935:
936: \subsection{Strategy} \label{s4.1}
937:
938: To bring out the similarities and differences between the reduced
939: model and the full theory, it is instructive to begin with the
940: Hamiltonian constraint (\ref{fullham}) of the full theory:
941: %
942: \begin{equation}
943: \ham= \int\md^3 x N e^{-1}\, [\epsilon_{ijk} E^{ai} E^{bj}\,
944: F_{ab}^k- 2(1+\gamma^2) E_i^aE_j^b\, K_{[a}^iK_{b]}^j ] \, .
945: \end{equation}
946: %
947: We begin by noting some simplifications that arise because of
948: spatial homogeneity. Recall that the connection $A$ is related to
949: the spin-connection $\Gamma$ defined by the triad and the
950: extrinsic curvature $K$ via $A= \Gamma + \gamma K$. As remarked in
951: section \ref{s2.1}, symmetries of the model imply that
952: $\Gamma$ is the standard magnetic monopole
953: connection; it does not depend on the phase space point under
954: consideration. From its expression (\ref{Gamma}), it follows that
955: its curvature $\Omega$ is given by:
956: %
957: \begin{equation}
958: \Omega = -\sin\vt\,\tau_3\, \md\vt \wedge \md\vp
959: \end{equation}
960: %
961: Since it is a `c-number' it can be trivially taken over to quantum
962: theory. Now, the curvature $F$ of the full connection $A$ can be
963: expanded out as:
964: %
965: \be F_{ab} = 2\partial_{[a}A_{b]} + [A_a,\, A_b] = \Omega_{ab} +
966: 2\gamma\partial_{[a}K_{b]} + \gamma^2 [K_a,\, K_b] + \gamma [\Gamma_a,\,
967: K_b]- \gamma [\Gamma_b,\, K_a]\ee
968: %
969: Using the expressions (\ref{Gamma}), (\ref{K}) and (\ref{sym2}) of
970: $\Gamma$, $K$ and $E$, it is straightforward to verify that:
971: %
972: \ba
973: [\Gamma, \, \gamma K ] = b \tau_2 \cos \vt \md\vt\wedge \md\vp
974: &=& -\gamma \md K\nonumber\\
975: \epsilon_{ijk} (\partial_{[a} K_{b]}^i) \,E^a_j E^b_k &=& 0
976: \label{simplify} \ea
977: %
978: These equations now imply that the integrand of the Hamiltonian
979: constraint can be written simply as%
980: %
981: \footnote{This simplification has a natural origin. Because of
982: (\ref{simplify}), the right hand side is just
983: $\epsilon_{ijk}E^{ai} E^{bj}\, {}^+\!F_{ab}^k$, where ${}^+\!F$ is
984: the curvature of the self-dual connection ${}^+\!A = \Gamma + i
985: K$.}
986: %
987: %
988: \be \ham(x) = e^{-1}\,E^{ai} E^{bj}\, (\epsilon_{ijk} F_{ab}^k-
989: 2(1+\gamma^2) K_{[a}^iK_{b]}^j) = e^{-1}\,E^{ai} E^{bj}\,
990: (\epsilon_{ijk}\Omega_{ab}^k - 2 K_{[ai} K_{b]j})\, . \ee
991: %
992: Since $\Omega$ is constant on the entire phase space, the
993: non-trivial part of the Hamiltonian constraint is thus contained
994: just in a term quadratic in extrinsic curvature.
995:
996: To pass to the quantum theory, we have to express the right hand
997: side in terms of holonomies and triads. There are two possible
998: avenues.\\
999: i) Using (\ref{simplify}), the extrinsic curvature terms in the
1000: constraint function $\ham(x)$ can be written in terms of
1001: curvatures $F$ and $\Omega$ so that we have:
1002: %
1003: \be \label{hami}
1004: \ham(x) = \frac{1}{\gamma^2}\, \epsilon_{ijk}\, e^{-1}\,
1005: E^{ai}E^{bj}\, [(1 +\gamma^2)\, \Omega_{ab}^k - F_{ab}^k]\, .
1006: \ee
1007: %
1008: The idea now is to use the discussion of co-triads in section
1009: \ref{s3.2} to obtain the operator analog of $\epsilon_{ijk} \,
1010: e^{-1}\,E^{ai}E^{bj}$; carry $\Omega$ to quantum theory trivially
1011: and use holonomies to express the field strength $F$ as in the
1012: full theory.\\
1013: ii) Alternatively, since we have completely fixed the gauge
1014: freedom to perform internal $\SU(2)$ rotations, we can regard $K$
1015: itself as a connection. It is a direct analog of the connection
1016: $A$ in the spatially flat cosmologies where $\Gamma$ vanishes.
1017: Therefore, we will denote its curvature by $\F$; $\F = \md K + [K,\,
1018: K]$. Then, the constraint functional can also be expressed as:
1019: %
1020: \be \label{hamii}
1021: \ham(x) = \frac{1}{\gamma^2}\, \epsilon_{ijk}\,e^{-1}\,
1022: E^{ai}E^{bj}\, [\gamma^2\, \Omega_{ab}^k - \F_{ab}^k] \ee
1023: %
1024: Now, one can again use the discussion of co-triads in section
1025: \ref{s3.2} to obtain the operator analog of $\epsilon_{ijk}\,
1026: e^{-1}\,E^{ai}E^{bj}$; carry $\Omega$ to quantum theory trivially
1027: and use holonomies to express the field strength $\F$. However,
1028: now holonomies have to be constructed using the connection $K$
1029: rather than $A$. Since our Hilbert space $\H$ is built from
1030: holonomies of $A$, at first it seems difficult to express
1031: holonomies of $K$ as operators on $\H$. However, because of
1032: spatial homogeneity, we can calculate $\ham(x)$ at any point. Let
1033: us choose a point on the equator and use holonomies along the
1034: three sets of curves introduced in (\ref{hol1})--(\ref{hol3}).
1035: Using the expressions (\ref{sym1}) and $(\ref{K})$ of $A$ and
1036: $K$, it follows that, along these curves the holonomies of $A$ and
1037: $K$ are equal! Therefore, the required holonomies of $K$ can
1038: indeed be expressed as operators on $\H$.
1039:
1040: Both these strategies are viable and, since they begin with
1041: \emph{exact} expressions (\ref{hami}) and (\ref{hamii}) of
1042: classical constraints, the difference in the resulting operators
1043: is just a quantization ambiguity. The first strategy is more
1044: natural in the sense that it does not require the introduction of
1045: a new connection. On the other hand, since the new connection $K$
1046: is the direct analog of the connection used in spatially flat
1047: cosmologies (in particular in \cite{abl}), to facilitate
1048: comparisons, the second strategy has been used in general,
1049: homogeneous models \cite{mb6,bdv}. For definiteness, we will use
1050: the second strategy in the main discussion and comment at the end
1051: on how the first strategy modifies the final constraint
1052: operator. The analysis of the resolution of the singularity can be
1053: carried out with either methods and the conclusion is the same.
1054:
1055:
1056: \subsection{Quantum Hamiltonian Constraint} \label{s4.2}
1057:
1058: Starting from the classical Hamiltonian constraint
1059: %
1060: \be \label{ham} \ham = \frac{1}{\gamma^2}\, \int \md^3 x\,\,
1061: \epsilon_{ijk}\, e^{-1}\, E^{ai}E^{bj}\, [\gamma^2\, \Omega_{ab}^k
1062: - \F_{ab}^k]\, ,\ee
1063: %
1064: where the integral is taken over the elementary cell, we wish to
1065: pass to the quantum operator. The overall strategy is the same as
1066: in \cite{abl,mb6,bdv} and the subtleties are discussed in detail
1067: in \cite{abl}. Therefore, here we will present only the main steps
1068: and comment on differences from the isotropic case treated in
1069: \cite{abl}.
1070:
1071: In the isotropic case, the elementary cell is a cube of length
1072: $L_o$ with respect to the fiducial metric. To express the
1073: curvature of the connection and the co-triad operator in the
1074: Hamiltonian constraint, one uses holonomies along the curves of
1075: variable length $\mu_oL_o$ along the edges of this cell. For each
1076: value of $\mu_o$ one obtains a quantum constraint operator
1077: $\hat{{\cal C}}_{\rm Ham}^{(\mu_o)}$. As discussed in \cite{abl},
1078: considerations from the full theory imply that $\mu_o$ should not
1079: be regarded as a regulator; rather, the $\mu_o$-dependence of this
1080: operator should be regarded as a quantization ambiguity. This
1081: ambiguity could not be fixed within the reduced model itself. But
1082: by making an appeal to results from the full theory, one can fix
1083: the value of $\mu_o$ using the minimum eigenvalue of the area
1084: operator in the full theory. (Recall that area enclosed by a loop
1085: enters while expressing curvature in terms of holonomies.)
1086:
1087: In the present, Kantowski-Sachs model, the overall procedure is
1088: the same but minor adjustments are necessary because of lack of
1089: isotropy. Since $\S^2$ is compact, our elementary cell $\S^2\times
1090: {\cal I}$ has a geometric edge only along the ${\cal I}$
1091: direction. As in section \ref{s3.1}, we will supplement it with
1092: edges along a longitude and the equator of $\S^2$. Holonomies
1093: along these three edges are then given by
1094: (\ref{hol1})--(\ref{hol3}). Now consider curves of length $L_o\delta$
1095: along ${\cal I}$ and of length $\delta$ each along the equator and the
1096: longitude of $\S^2$. Then, the co-triad function in the
1097: Hamiltonian (\ref{ham}) can be expressed as:
1098: %
1099: \begin{equation}\label{cotraid}
1100: \epsilon_{ijk}\tau^i\, e^{-1}E^{aj}E^{bk}= -2(8\pi\gamma G{\cal
1101: L}_{(k)})^{-1}\, \epsilon^{abc}\;\,\w_c^k\,
1102: h_k^{(\delta)}\{h_k^{(\delta)-1},V\}
1103: \end{equation}
1104: %
1105: where on the right hand side the index $k$ runs over $x,\vt,\vp$
1106: and is summed over; $h_k$ is the holonomy along the edge $k$; and
1107: ${\cal L}_x = L_o\delta$ and ${\cal L}_\vt = {\cal L}_\vp = \delta$. Note
1108: that this is an exact equality for any choice of $\delta$. Components
1109: of the curvature $\F$ can also be expressed in terms of these
1110: holonomies. Set
1111: %
1112: \begin{equation}
1113: h^{(\delta)}_{ij} = h_i^{(\delta)}h_j^{(\delta)}(h_i^{(\delta)})^{-1}
1114: (h_j^{(\delta)})^{-1}
1115: \end{equation}
1116: %
1117: where $i,j,k$ run over $x,\vt,\vp$. Then, using
1118: (\ref{hol1})--(\ref{hol3}) it is straightforward to verify that
1119: %
1120: \begin{equation} \label{F}
1121: \F_{ab}^i(x)\tau_i=\frac{\w^i_a\w^j_b}{{\cal A}_{(ij)}}\,\,
1122: (h^{{\cal A}_{(ij)}}_{ij}-1)\, +\,\,O((b^2+c^2)^{3/2}\sqrt{\cal
1123: A})
1124: \end{equation}
1125: %
1126: where ${\cal A}_{x\vt}=\delta^2L_o={\cal A}_{x\vp}$ and ${\cal
1127: A}_{\vt\vp}=\delta^2$. For $\F_{x,\vt}$ and $\F_{x,\vp}$, (\ref{F}) is
1128: the standard geometric relation between holonomies around closed
1129: loops, the area they enclose and curvature. For $\F_{\vt,\vp}$, on
1130: the other hand, while (\ref{F}) continues to hold, the standard
1131: geometrical interpretation is no longer available because the
1132: edges of length $\delta$ along the equator and a longitude fail to
1133: form a closed loop. Nonetheless, the standard relation is
1134: meaningfully extended because of spatial homogeneity, $\delta^2$
1135: playing the role of the `effective area' ${\cal A}_{\vt,\vp}$.
1136:
1137: We now have all the ingredients, $\F$, $\Omega$ and the triad
1138: components, to rewrite the classical constraint (\ref{ham}) in a
1139: way which is suitable for quantization. For the $\F$-term in
1140: (\ref{ham}) we obtain
1141: %
1142: \begin{eqnarray}\label{Fterm}
1143: &-&\gamma^{-2}\int\md^3x\, \epsilon_{ijk}\, \F_{ab}^i\,\,
1144: e^{-1}E^{aj}E^{bk}
1145: = 2\gamma^{-2}\int\md^3x\,\, \tr\left( \F_{ab}^i\tau_i\;
1146: \epsilon_{jkl}\, e^{-1} E^{aj}E^{bk}\tau^l\right)\nonumber\\
1147: &=& -4(8\pi\gamma^3G)^{-1}\int\md^3x
1148: \det\w\;\tr\left(\epsilon^{ijk}{\cal A}_{(ij)}^{-1}
1149: (h_{ij}^{(\delta)}-1){\cal L}_{(k)}^{-1}h_k^{(\delta)}
1150: \{h_k^{(\delta)-1},V\}\right) \,+O((b^2+c^2)^{3/2}\delta) \nonumber\\
1151: &=& \left[-16\pi(8\pi\gamma^3G\delta^3)^{-1}\sum_{ijk}\epsilon^{ijk}
1152: \tr\left(h_{ij}^{(\delta)}h_k^{(\delta)} \{h_k^{(\delta)},V\}\right)\right]\,
1153: +O\left((b^2+c^2)^{3/2}\delta\right)\, .
1154: \end{eqnarray}
1155: %
1156: Next, using $\Omega = -\sin\vt\md\vt\wedge\md\vp \tau_3$, the
1157: $\Omega$ term in (\ref{ham}) becomes:
1158: %
1159: \begin{eqnarray}\label{Omegaterm}
1160: \int\md^3x\,\epsilon_{ijk}\, \Omega_{ab}^i\,\,
1161: e^{-1}E^{aj}E^{bk}\,\, \; &=& -2\int\md^3x\,
1162: \tr\left(\Omega_{ab}^i\tau_i\;
1163: \epsilon_{jkl}\, e^{-1}E^{al}E^{bk}\tau^l\right)\\
1164: &=& -8(8\pi\gamma G\delta)^{-1}\int\md^3x\sin\vt\;
1165: \tr\left(\tau_3\;{\cal L}_x^{-1}h_x^{(\delta)}\{h_x^{(\delta)-1},V\}\right)\\
1166: &=& -32\pi(8\pi\gamma G\delta)^{-1}\;
1167: \tr\left(\tau_3\;h_x^{(\delta)}\{h_x^{(\delta)-1},V\}\right)\,.
1168: \end{eqnarray}
1169: %
1170: Set
1171: %
1172: \be C^{(\delta)} = \big[-2 (\gamma^3
1173: G\delta^3)^{-1}\,\sum_{ijk}\epsilon^{ijk}
1174: \tr\left(h_{ij}^{(\delta)}h_k^{(\delta)} \{h_k^{(\delta)},V\}\right)\big] -
1175: \big[-4(\gamma G \delta)^{-1}
1176: \tr\left(\tau_3\;h_x^{(\delta)}\{h_x^{(\delta)-1},V\}\right)\, \big]\ee
1177: %
1178: Then the classical Hamiltonian constraint is given by
1179: %
1180: \be \ham = \lim_{\delta \to 0}\, C^{(\delta)} \ee
1181: %
1182: Since all terms in $C^{\delta}$ are expressed purely in terms of our
1183: elementary variables ---holonomies and triads--- which have direct
1184: operator analogs, passage to quantum theory is now
1185: straightforward. We obtain:
1186: %
1187: \begin{eqnarray}
1188: \hat{C}^{(\delta)} &=& 2i
1189: (\gamma^3\delta^3\lp^2)^{-1} \tr\left(\sum_{ijk}
1190: \epsilon^{ijk}\hat{h}_i^{(\delta)} \hat{h}_j^{(\delta)} \hat{h}_i^{(\delta)-1}
1191: \hat{h}_j^{(\delta)-1}\hat{h}_k^{(\delta)}
1192: [\hat{h}_k^{(\delta)-1},\hat{V}]+2\gamma^2\delta^2\tau_3 \hat{h}_x^{(\delta)}
1193: [\hat{h}_x^{(\delta)-1},\hat{V}]\right)\nonumber\\\nonumber &=&
1194: 4i(\gamma^3\delta^3\lp^2)^{-1}\left(
1195: 8\sin\frac{\delta b}{2}\cos\frac{\delta b}{2} \sin\frac{\delta
1196: c}{2}\cos\frac{\delta c}{2}
1197: \left(\sin\frac{\delta b}{2}\hat{V}\cos\frac{\delta b}{2}-
1198: \cos\frac{\delta b}{2}\hat{V}\sin\frac{\delta b}{2}\right)\right.\\
1199: && + \left.\left(4\sin^2\frac{\delta b}{2}\cos^2\frac{\delta
1200: b}{2}+\gamma^2\delta^2\right)
1201: \left(\sin\frac{\delta c}{2}\hat{V}\cos\frac{\delta c}{2}-
1202: \cos\frac{\delta c}{2}\hat{V}\sin\frac{\delta c}{2}\right)\right)\label{C}
1203: \end{eqnarray}
1204: %
1205: (This is a special case of Eq. (26) in \cite{bdv}.)
1206:
1207: The action of this operator on the eigenstates
1208: $|\mu,\nu\rangle$ of $\hat{p}_b$ and $\hat{p}_c$ is given by
1209: %
1210: \begin{eqnarray}
1211: \hat{C}^{(\delta)} |\mu,\nu\rangle &=&
1212: (2\gamma^3\delta^3\lp^2)^{-1}
1213: \left[ 2(V_{\mu+\delta,\nu}-V_{\mu-\delta,\nu})\right.\\
1214: &&\times(|\mu+2\delta,\nu+2\delta\rangle- |\mu+2\delta,\nu-2\delta\rangle-
1215: |\mu-2\delta,\nu+2\delta\rangle+ |\mu-2\delta,\nu-2\delta\rangle)\nonumber\\
1216: &&+\left.(V_{\mu,\nu+\delta}-V_{\mu,\nu-\delta}) (|\mu+4\delta,\nu\rangle-
1217: 2(1+2\gamma^2\delta^2)|\mu,\nu\rangle+ |\mu-4\delta,\nu\rangle)\right]\,.
1218: \nonumber
1219: \end{eqnarray}
1220: %
1221:
1222: However, while the classical constraint is a real function on the
1223: kinematical phase space ${\bf \Gamma}$, $\hat{C}^{\delta}$ fails to be
1224: self-adjoint on the kinematical Hilbert space $\H$. We therefore
1225: need to add to it its Hermitian adjoint.%
1226: %
1227: \footnote{Although the self-adjoint form of the Hamiltonian
1228: constraint was discussed briefly in \cite{mb7}, this point was
1229: ignored in the detailed treatment of \cite{abl}. A detailed
1230: treatment of the self-adjoint constraint and its semi-classical
1231: implications can be found in \cite{jw,abw}. It is used in a
1232: crucial way to obtain the physical Hilbert space in \cite{aps}.}
1233: %
1234: The result is an operator $\hat{C}_{\rm grav}^{(\delta)}:=\frac{1}{2}
1235: (\hat{C}^{(\delta)}+\hat{C}^{(\delta)\dagger})$ given by:
1236: %
1237: \begin{eqnarray}
1238: \hat{C}_{\rm grav}^{(\delta)} |\mu,\nu\rangle &=&
1239: (2 \gamma^3\delta^3\lp^2)^{-1}\,
1240: \left[(V_{\mu+\delta,\nu}-V_{\mu-\delta,\nu}+V_{\mu+3\delta,\nu+2\delta}-
1241: V_{\mu+\delta,\nu+2\delta}) |\mu+2\delta,\nu+2\delta\rangle \right.\nonumber\\
1242: &&- (V_{\mu+\delta,\nu}-V_{\mu-\delta,\nu}+V_{\mu+3\delta,\nu-2\delta}-
1243: V_{\mu+\delta,\nu-2\delta}) |\mu+2\delta,\nu-2\delta\rangle \nonumber\\
1244: &&-(V_{\mu+\delta,\nu}-V_{\mu-\delta,\nu}+V_{\mu-\delta,\nu+2\delta}-
1245: V_{\mu-3\delta,\nu+2\delta}) |\mu-2\delta,\nu+2\delta\rangle \nonumber\\
1246: &&+ (V_{\mu+\delta,\nu}-V_{\mu-\delta,\nu}+V_{\mu-\delta,\nu-2\delta}-
1247: V_{\mu-3\delta,\nu-2\delta}) |\mu-2\delta,\nu-2\delta\rangle \nonumber\\
1248: &&+{\textstyle\frac{1}{2}}(V_{\mu,\nu+\delta}-V_{\mu,\nu-\delta}+V_{\mu+4\delta,\nu+\delta}-
1249: V_{\mu+4\delta,\nu-\delta}) |\mu+4\delta,\nu\rangle \nonumber\\
1250: &&-(1+2\gamma^2\delta^2)(V_{\mu,\nu+\delta}-V_{\mu,\nu-\delta}) |\mu,\nu\rangle
1251: \nonumber\\
1252: &&+\left.{\textstyle\frac{1}{2}}
1253: (V_{\mu,\nu+\delta}-V_{\mu,\nu-\delta}+V_{\mu-4\delta,\nu+\delta}-
1254: V_{\mu-4\delta,\nu-\delta}) |\mu-4\delta,\nu\rangle\right]\,.
1255: \end{eqnarray}
1256:
1257: Physical states in quantum theory are those which are symmetric
1258: under the `parity operator' $\hat{\Pi}_b$ and lie in the kernel of
1259: the operator $\hat{C}_{\rm grav}^{(\delta)}$. That is, the only
1260: non-trivial quantum Einstein's equations are:
1261: %
1262: \be \label{qee1} (\Psi| \hat{\Pi}_b = 0 \quad,\quad (\Psi|
1263: \hat{C}_{\rm grav}^{(\delta)} = 0 \ee
1264: %
1265: where $(\Psi|$ is an element of the dual $\cylstar$ of the space
1266: $\cyl$ of finite linear combinations of almost periodic functions
1267: of $b,c$. Let us expand $(\Psi|$ using eigenbras $(\mu,\nu|$ (in
1268: $\cylstar$) of the triad operators:
1269: %
1270: \be (\Psi| = \sum_{\mu,\nu} \, \psi_{\mu,\nu}\, (\mu,\nu| \ee
1271: %
1272: Then, $\psi_{\mu,\nu}$ can be regarded as wave functions in the
1273: triad (or, Riemannian geometry) representation. To exhibit the
1274: action of $\hat{C}_{\rm grav}^{(\delta)}$ on $\psi_{\mu,\nu}$ we are
1275: led to separate the cases $\mu \ge 4\delta$ from $\mu < 4\delta$
1276: because of the absolute-value $|\mu|$ in the volume eigenvalues.
1277: For $\mu \ge 4\delta$, quantum Einstein's equation (\ref{qee1})
1278: becomes
1279: %
1280: \begin{eqnarray} \label{qee2}
1281: \hat{C}_{\rm grav}^{(\delta)}\, \psi_{\mu,\nu}
1282: &=&2\delta(\sqrt{|\nu+2\delta|}+\sqrt{|\nu|})
1283: \left(\psi_{\mu+2\delta,\nu+2\delta}- \psi_{\mu-2\delta,\nu+2\delta}\right)
1284: \nonumber\\
1285: && +(\sqrt{|\nu+\delta|}-\sqrt{|\nu-\delta|})
1286: \left((\mu+2\delta)\psi_{\mu+4\delta,\nu}-
1287: (1+2\gamma^2\delta^2)\mu\psi_{\mu,\nu}+
1288: (\mu-2\delta)\psi_{\mu-4\delta,\nu}\right)\nonumber\\
1289: &&+2\delta(\sqrt{|\nu-2\delta|}+\sqrt{|\nu|})
1290: \left(\psi_{\mu-2\delta,\nu-2\delta}-
1291: \psi_{\mu+2\delta,\nu-2\delta}\right)\nonumber\\
1292: &=& 0\, .
1293: \end{eqnarray}
1294: %
1295: For $\mu=0$, we have
1296: %
1297: \begin{equation} \label{qee3mu0}
1298: \sqrt{|\nu+2\delta|} \psi_{2\delta,\nu+2\delta}
1299: +(\sqrt{|\nu+\delta|}-\sqrt{|\nu-\delta|})
1300: \psi_{4\delta,\nu}
1301: -\sqrt{|\nu-2\delta|}\psi_{2\delta,\nu-2\delta}= 0\, ;
1302: \end{equation}
1303: %
1304: for $\mu=\delta$:
1305: %
1306: \begin{eqnarray} \label{qee3mu1}
1307: &&2(\sqrt{|\nu+2\delta|}+\sqrt{|\nu|}) \psi_{3\delta,\nu+2\delta}+
1308: 2(\sqrt{|\nu+2\delta|}-\sqrt{|\nu|}) \psi_{\delta,\nu+2\delta}\nonumber\\
1309: && +(\sqrt{|\nu+\delta|}-\sqrt{|\nu-\delta|})
1310: \left(3\psi_{5\delta,\nu}+2\psi_{3\delta,\nu}- (1+2\gamma^2\delta^2)\psi_{\delta,\nu}
1311: \right)\nonumber\\
1312: &&-2(\sqrt{|\nu-2\delta|}+\sqrt{|\nu|})\psi_{3\delta,\nu-2\delta}-
1313: 2(\sqrt{|\nu-2\delta|}-\sqrt{|\nu|})\psi_{\delta,\nu-2\delta}
1314: = 0\,,
1315: \end{eqnarray}
1316: %
1317: for $\mu=2\delta$:
1318: %
1319: \begin{eqnarray} \label{qee3mu2}
1320: &&(\sqrt{|\nu+2\delta|}+\sqrt{|\nu|}) \psi_{4\delta,\nu+2\delta}-
1321: \sqrt{|\nu|}\psi_{0,\nu+2\delta}\nonumber\\
1322: && +(\sqrt{|\nu+\delta|}-\sqrt{|\nu-\delta|})
1323: \left(2\psi_{6\delta,\nu}+\psi_{2\delta,\nu}- (1+2\gamma^2\delta^2)\psi_{2\delta,\nu}
1324: \right)\nonumber\\
1325: &&-(\sqrt{|\nu-2\delta|}+\sqrt{|\nu|})\psi_{4\delta,\nu-2\delta}
1326: +\sqrt{|\nu|}\psi_{0,\nu-2\delta}=0\,,
1327: \end{eqnarray}
1328: %
1329: and for $\mu=3\delta$:
1330: %
1331: \begin{eqnarray} \label{qee3mu3}
1332: &&2(\sqrt{|\nu+2\delta|}+\sqrt{|\nu|}) (\psi_{5\delta,\nu+2\delta}-
1333: \psi_{\delta,\nu+2\delta})\nonumber\\
1334: && +(\sqrt{|\nu+\delta|}-\sqrt{|\nu-\delta|})
1335: \left(5\psi_{7\delta,\nu}+2\psi_{\delta,\nu}- 3(1+2\gamma^2\delta^2)\psi_{3\delta,\nu}
1336: \right)\nonumber\\
1337: &&-2(\sqrt{|\nu-2\delta|}+\sqrt{|\nu|})(\psi_{5\delta,\nu-2\delta}-
1338: \psi_{\delta,\nu-2\delta})=0\,.
1339: \end{eqnarray}
1340: %
1341: For values of $\mu$ not an integer multiple of $\delta$ one can derive
1342: similar expressions, but they will not be used in what follows. As
1343: remarked above, strictly speaking, the role of (\ref{qee2}) is
1344: only to select physical wave functions. However, intuitively it
1345: can also be thought of as the quantum evolution equation. Recall
1346: that $p_c$ can be regarded as an internal time in the classical
1347: theory of the model. Therefore, in the quantum theory, one may
1348: regard $\nu$ as internal time. Then, (\ref{qee2}) can be
1349: interpreted as the quantum Einstein's equation, which evolves the
1350: wave functions in discrete steps of magnitude $2\delta$ along $\nu$.
1351:
1352: So far, the parameter $\delta$ ---and hence the size of the
1353: `time-step' --- is arbitrary. In the classical theory, we recover
1354: the Hamiltonian constraint only in the limit $\delta$ goes to
1355: zero. This is because, while the expression (\ref{cotriad}) is
1356: exact for all loops, curvature ${}^oF$ can be expressed in terms
1357: of holonomies only in the limit in which the areas of loops shrink
1358: to zero (see (\ref{F})). In quantum theory, however, the
1359: straightforward limit $\lim_{\delta\to 0} \hat{C}^{(\delta)}$
1360: diverges (for the same reasons as in the isotropic case
1361: \cite{abl}). The viewpoint is that this is occurs because the
1362: limit ignores the quantum nature of geometry, i.e., the fact that
1363: in full quantum geometry, the area operator has a minimum non-zero
1364: eigenvalue. The `correct' quantization of the Hamiltonian
1365: constraint $\ham$ has to take in to account the quantum nature of
1366: geometry.
1367:
1368: To do so, note first that the holonomies used in the expression of
1369: $\F$ define quantum states with $\mu=\delta$ and $\nu=\delta$. Using these
1370: states, we can calculate the \emph{quantum} area of each of the
1371: three faces of the elementary cell $\S^2\times {\cal I}$, enclosed
1372: by the three curves ---one along $x$, one along a longitude and
1373: one along the equator of $\S^2$. The area operators defined by
1374: faces $S_{x,\vt}, S_{x,\vp}$ and $S_{\vt,\vp}$ are, respectively,
1375: $2\pi |\hat{p}_b|,\, 2\pi |\hat{p}_b|$ and $\pi |\hat{p}_c|$ (see
1376: (\ref{area})). The quantum geometry states defined by the three
1377: holonomies are eigenstates of these area operators. Furthermore,
1378: they have the same eigenvalue: $\pi\gamma\delta \lp^2$. Now, in the
1379: full theory, we know that the area operator has a minimum non-zero
1380: eigenvalue, $a_o = 2\sqrt{3}\pi\gamma \lp^2$. The viewpoint, as in
1381: \cite{abl}, is that in the calculation of the field strength $\F$,
1382: it is \emph{physically} inappropriate to try to use surfaces of
1383: arbitrarily small areas. The best one can do is to shrink the area
1384: of the loop till it attains this quantum minimum $a_o$. This
1385: implies that we should set $\delta = 2\sqrt{3}$. To summarize, by
1386: using input from the full theory, we conclude that the quantum
1387: Hamiltonian constraint is given by $\hat{C}^{(\delta_o)}$ with $\delta_o =
1388: 2\sqrt{3}$.
1389:
1390:
1391: \emph{Remarks}:
1392:
1393: i) Note that the character of the difference equation changes
1394: depending on whether $\mu\ge 4\delta$ or $\mu < 4\delta$. In particular,
1395: for $\mu \ge 4\delta$, the knowledge of the wave function at `times'
1396: $\nu + 2\delta$ and $\nu$ determines only the combination
1397: $\psi_{\mu-2\delta, \nu-2\delta} - \psi_{\mu +2\delta,\nu-2\delta}$ at `time'
1398: $\nu-2\delta$ via (\ref{qee2}). On the other hand, for $\mu =0$ the
1399: wave function at $\nu + 2\delta$ and $\nu$ determines $\psi_{2\delta,
1400: \tau+2\delta}$ completely via (\ref{qee3mu0}). We will return to these
1401: differences in section \ref{s4.3}.
1402:
1403: ii) As discussed in \cite{abl,mb8,bd} in detail, one can recover
1404: the Wheeler-DeWitt equation with a specific factor ordering by
1405: taking a systematic limit of quantum Einstein's equation. In the
1406: present model, the difference equation reduces to the following
1407: differential equation for the wave function
1408: $\Psi(p_b,p_c)=\psi_{2p_b/\gamma\lp^2, \, p_c/\gamma\lp^2}$ on the
1409: classical minisuperspace $\M$:
1410: %
1411: \be\label{wdw} {16\lp^4}\,\left( \sqrt{p_c}\,\f{\partial^2
1412: \Psi}{\partial p_b \partial p_c} + \f{p_b}{4\sqrt{p_c}}\,
1413: \f{\partial^2 \Psi}{\partial p_b^2}+ \frac{1}{2\sqrt{p_c}}
1414: \frac{\partial\Psi}{\partial p_b} \right) - 4 \f{p_b}{\sqrt{p_c}}
1415: \Psi =0 \ee
1416: %
1417: For $\mu \gg \delta, \nu\gg \delta$, solutions of the discrete equation
1418: (\ref{qee2}) can be approximated by those of this Wheeler DeWitt
1419: equation in a precise sense. In essence, this Wheeler-DeWitt
1420: equation is obtained from (\ref{qee2}) by ignoring quantum
1421: geometry effects, i.e., by an appropriate $\delta \to 0$ limit. For
1422: the approximation by a differential equation, higher derivatives
1423: of the wave function $\psi$ must be sufficiently small. Thus, the
1424: limit $\delta\to0$ for the equation (\ref{qee2}) exists only under
1425: additional assumptions on the wave function. It does not exist for
1426: the \emph{operator} $\hat{C}_{\rm grav}^{(\delta)}$ by itself.
1427:
1428: \subsection{Absence of Singularity}
1429: \label{s4.3}
1430:
1431: Recall that the classical singularity occurs at $p_c =0$. The
1432: Wheeler DeWitt equation (\ref{wdw}) is manifestly singular there.
1433: One can multiply it by $\sqrt{p_c}$ and define a new `internal
1434: time, $\bar{p_c} = \ln p_c$ and obtain a regular equation in
1435: variables $p_b,\bar{p}_c$. However, since $\bar{p}_c \rightarrow
1436: -\infty$ at the singularity, this regular equation does not let us
1437: evolve across the singularity. Thus, the overall situation is the
1438: same as in the isotropic model.
1439:
1440: To obtain the Wheeler-DeWitt equation from the `fundamental'
1441: difference equation (\ref{qee2}), we had to ignore quantum
1442: geometry effects. Thus, the key question is whether this failure
1443: of the Wheeler-DeWitt equation is an artifact of the approximation
1444: used. Can quantum geometry effects make a qualitative difference
1445: as they do in the isotropic model? We will now argue that the
1446: answer is in the affirmative.
1447:
1448: Let us analyze the `evolution' given by the difference equation
1449: (\ref{qee2}). Does this evolution stop at $\nu =0$? As in the
1450: isotropic case \cite{mb3,mb7}, we will use the quantum constraint
1451: (\ref{qee2}) as a recurrence relation starting at positive $\nu$
1452: and evolve toward smaller values. However, now a new twist arises
1453: because for generic values $\mu \ge 4\delta$, (\ref{qee2}) determines
1454: only the difference $\psi_{\mu+2\delta,\nu-2\delta}
1455: -\psi_{\mu-2\delta,\nu-2\delta}$ as a function of the initial values of
1456: the wave function. Therefore, quantum Einstein's equation has to
1457: be supplemented with an appropriate boundary condition. From
1458: (\ref{qee2}) which holds for $\mu \ge 4\delta$, one might conclude
1459: that, even if we restrict ourselves to the `lattice' $\mu = n\delta$,
1460: one would have to specify the wave function at $\mu=0, \delta, 2\delta,
1461: 3\delta$ at each `time step'. However, as remarked at the end of
1462: section \ref{s4.2}, the form of the difference equations for $\mu
1463: < 4\delta$ is different. An examination of (\ref{qee3mu0}) and
1464: (\ref{qee3mu1}) reveals that it is in fact sufficient to specify
1465: the wave function $\psi_{\mu,\nu}$ just at $\mu=0$ and $\mu=\delta$ at
1466: each `time-step'. This is a mathematically viable choice of the
1467: boundary condition and heuristically it corresponds to providing
1468: data the `horizon'. It will turn out that the issue of the
1469: resolution of singularity is insensitive to the precise choice of
1470: the boundary condition. We will therefore postpone the discussion
1471: of physically appropriate choices until after the main discussion.
1472:
1473: Let us then start at some finite positive $\nu$ and carry out a
1474: backward evolution using (\ref{qee2}). In contrast to the
1475: Wheeler-DeWitt equation (\ref{wdw}), the coefficients in the
1476: difference equation (\ref{qee2}) are always regular. At first one
1477: might think that this regularity by itself would be sufficient to
1478: ensure that the evolution would be well-defined across the
1479: singularity. Note however that this need not be the case. For, the
1480: backward evolution continues only as long as the coefficient of
1481: the wave function at $\nu-2\delta$ is non-zero. The key question
1482: therefore is whether this coefficient vanishes. The issue is
1483: subtle. For instance, the coefficients in the non-self-adjoint
1484: operator $(\hat{C}^{\delta})^\dag$ are also non-singular. However, if
1485: one uses it in place of $\hat{C}^{\delta}_{\rm Ham}$, one finds that
1486: the coefficient vanishes and one can not evolve across the
1487: singularity.
1488:
1489: For our quantum Einstein's equation (\ref{qee2}), this coefficient
1490: is given by $\sqrt{|\nu-2\delta|}+\sqrt{|\nu|}$. By inspection,
1491: \emph{it never vanishes}. Therefore, the backward quantum
1492: evolution remains well-defined and determines the wave function
1493: not only for $\nu >0$ but also in the new region with $\nu\le 0$.
1494: In this precise sense, the classical black hole singularity can be
1495: traversed using quantum evolution and thus ceases to be a boundary
1496: of space-time.
1497:
1498: Next, recall from section \ref{s4.1} that we could have begun with
1499: the expression (\ref{hami}) of the Hamiltonian constraint in the
1500: classical theory and then proceeded with quantization. What would
1501: be the status of the singularity resolution with this choice? The
1502: procedure to construct the quantum Hamiltonian constraint would be
1503: identical. However, the final result would be slightly different:
1504: $\gamma^2$ in (\ref{qee2}) would be replaced by $1+\gamma^2$. That
1505: is, in the new quantum Einstein equation, the coefficient of
1506: $\psi_\mu,\nu$ in (\ref{qee2}) would be modified and the equation
1507: would become:
1508: %
1509: \begin{eqnarray} \label{qee4}
1510: &&2\delta(\sqrt{|\nu+2\delta|}+\sqrt{|\nu|}) (\psi_{\mu+2\delta,\nu+2\delta}-
1511: \psi_{\mu-2\delta,\nu+2\delta})\nonumber\\
1512: && +(\sqrt{|\nu+\delta|}-\sqrt{|\nu-\delta|})
1513: \left((\mu+2\delta)\psi_{\mu+4\delta,\nu}- (1+2(\gamma^2+1)
1514: \delta^2)\mu\psi_{\mu,\nu}+
1515: (\mu-2\delta)\psi_{\mu-4\delta,\nu}\right)\nonumber\\
1516: &&+2\delta(\sqrt{|\nu-2\delta|}+\sqrt{|\nu|})(\psi_{\mu-2\delta,\nu-2\delta}-
1517: \psi_{\mu+2\delta,\nu-2\delta})\nonumber\\
1518: &=& 0
1519: \end{eqnarray}
1520: %
1521: (The same replacement holds for $\mu < 4\delta$.) Since the
1522: coefficient of $\psi_{\mu-2\delta,\nu-2\delta}- \psi_{\mu+2\delta,\nu-2\delta}$ is
1523: unaltered, one would again be able to `evolve' across the
1524: singularity. Thus, the conclusion is robust within quantum
1525: geometry.%
1526: %
1527: \footnote{However, the relation between the wave function and the
1528: Wheeler--DeWitt approximation is less direct with this form of the
1529: constraint.}
1530: %
1531:
1532: To conclude, we will briefly return to the question of the
1533: boundary condition that are needed to make the evolution
1534: well-defined even away from the singularity. As noted earlier, in
1535: the backward evolution given by (\ref{qee3mu0}),\,
1536: $\psi_{2\delta,\nu_o}$ is determined by values of the wave function at
1537: $\nu>\nu_o$, so long as for $\nu_o\not=0$. Furthermore,
1538: (\ref{qee2}) then determines the wave function at $\mu=(2+4n)\delta$
1539: for all $n\in{\mathbb N}$ and at that $\nu_o$. One could now use
1540: pre-classicality arguments \cite{mb4} (i.e.\ require that the wave
1541: function not oscillate on small scales) at large $\mu$ to choose
1542: the boundary values, eliminating entirely the need for specifying
1543: boundary conditions. While this could be reasonable in
1544: semiclassical regimes, for small $|\nu|$ it would be questionable
1545: to use pre-classicality in the $\mu$-direction even if a
1546: pre-classical solution exist (which is not guaranteed, see e.g.\
1547: the analysis in \cite{ck,gd}). More importantly, the slice $\nu=0$
1548: is special because $\psi_{2\delta,0}$ is not determined through
1549: (\ref{qee3mu0}) or otherwise. At this point, the condition
1550: (\ref{qee3mu0}) evaluated at $\nu=-2\delta$ gives instead a condition
1551: for previous values of the wave function, translating into one
1552: condition for the initial values. At $\nu=0$, however,
1553: $\psi_{2\delta,0}$ has to be specified as boundary value in addition
1554: to $\psi_{0,0}$ because, unlike similar situations in isotropic
1555: models, it does not drop out of the evolution: it is needed in the
1556: recurrence (\ref{qee2}) for $\mu=4\delta$, $\nu=2\delta$. There is thus no
1557: reduction in conditions on the wave function from the fact that
1558: $\psi_{2\delta,0}$ drops out of (\ref{qee3mu0}); a condition is simply
1559: transferred from boundary values to initial values. Note however
1560: that, even though the line $\nu=0$ (corresponding to the classical
1561: singularity) does show special behavior, evolution does not break
1562: down there; it is only the boundary value problem which changes
1563: character.
1564:
1565: While such a boundary value problem is mathematically
1566: well-defined, the resulting theory is not necessarily physically
1567: correct. For, in a physically interesting theory one would expect
1568: that, well away from the singularity, there should exist
1569: semi-classical solutions which are peaked on the classical
1570: trajectories. From Fig.~1 it is clear that the semi-classical
1571: solution peaked on the trajectory labelled by mass $m$ would be
1572: sharply peaked at $\nu= \pm 4m^2$ on the line $\mu=0$, whence
1573: semi-classical states peaked at different classical solutions will
1574: have quite different forms near $\mu=0$. Therefore, the theory
1575: obtained by simply fixing the wave function on (or near) $\mu=0$
1576: is not likely to admit a sufficiently rich semi-classical sector.
1577: Indeed, it is not clear that any boundary condition at (or near)
1578: $\mu=0$ will be physically viable. Rather, one may have to impose
1579: the boundary condition at $\mu=\infty$, e.g., by requiring that
1580: the wave function should vanish there. Indeed, the form of
1581: dynamical trajectories of Fig.~1 implies that \emph{every}
1582: semi-classical state would share this property. Moreover, the
1583: corresponding condition does hold in the closed isotropic model
1584: with a massless scalar field as source \cite{aps2}. As in that
1585: case, one would expect that these semi-classical states would span
1586: a dense subspace in the physical Hilbert space. If so, requiring
1587: that the wave functions vanish at $\mu=\infty$ would be physically
1588: justified. Heuristics suggest that this strategy is viable for the
1589: Wheeler-DeWitt equation (\ref{wdw}). However, detailed numerical
1590: analysis is necessary to establish that the difference equations
1591: (\ref{qee2}) or (\ref{qee4}) admits solutions satisfying this
1592: condition (or a suitable modification thereof) for all $\nu$ and
1593: that its imposition makes the evolution unambiguous.
1594:
1595: \section{Discussion}
1596: \label{s5}
1597:
1598: Results of the last two sections support a general scenario that
1599: has emerged from the analysis of singularities in quantum
1600: cosmology. It suggests that the classical singularity does not
1601: represent a final frontier; the \emph{physical} space-time does
1602: not end there. In the Planck regime, quantum fluctuations do
1603: indeed become so strong that the classical description breaks
1604: down. The space-time continuum of classical general relativity is
1605: replaced by discrete quantum geometry which remains regular during
1606: the transition through what was a classical singularity. Certain
1607: similarities between the Kantowski-Sachs model analyzed here and a
1608: cosmological model which has been studied in detail \cite{aps}
1609: suggest that there would be a quantum bounce to another large
1610: classical region. If this is borne out by detailed numerical
1611: calculations, one would conclude that quantum geometry in the
1612: Planck regime serves as a bridge between two large classical
1613: regions. Space-time may be much larger than general relativity has
1614: had us believe.
1615:
1616: However, as indicated at the end of section \ref{s4.2} significant
1617: numerical work is still needed before one can be certain that this
1618: scenario is really borne out in the model. Moreover, this is a
1619: highly simplified model. It is important to check if the
1620: qualitative conclusions remain robust as more and more realistic
1621: features are introduced. First, one should extend the analysis so
1622: that the space-time region outside the horizon is also covered. A
1623: second and much more important challenge is incorporation of an
1624: infinite number of degrees of freedom by coupling the model, e.g.,
1625: to a spherically symmetric scalar field. First steps along these
1626: lines have been taken \cite{mbss,hw} and one can see that the
1627: evolution still extends beyond the classical singularity
1628: \cite{mb9}. But a comprehensive treatment still remains a distant
1629: goal.
1630:
1631: \section*{Acknowledgements:}
1632:
1633: This work was supported in part by NSF grants PHY-0090091, and
1634: PHY-0354932, the Alexander von Humboldt Foundation, the C.V. Raman
1635: Chair of the Indian Academy of Sciences and the Eberly research
1636: funds of Penn State.
1637:
1638: \begin{thebibliography}{99}
1639:
1640: \bibitem{cm} Misner C W 1972 Minisuperspaces, in \textit{Magic
1641: without magic, John A. Wheeler festschrift} ed Klauder J R
1642: (Freeman, San Fransisco)
1643:
1644: \bibitem{jh-lh} Hartle J B 1995 {Spacetime quantum mechanics and
1645: the quantum mechanics of spacetime}, in \textit{Gravitation and
1646: quantizations: Proceedings of the 1992 Les Houches summer school}
1647: (North Holland, Amsterdam) \textit{Preprint} \texttt{gr-qc/9304006}
1648:
1649: \bibitem{hh} Hartle J B and Hawking S W 1983 {The wave
1650: function of the universe} \textit{Phys.\ Rev.} \textbf{D28} 2960-2975
1651:
1652: \bibitem{alrev} Ashtekar A and Lewandowski L 2004 Background
1653: independent quantum gravity: A status report \textit{Class.
1654: Quant. Grav.} \textbf{21} R53-R152
1655:
1656: \bibitem{crbook} Rovelli C 2004 \emph{Quantum Ggravity} (Cambridge
1657: University Press, Cambridge)
1658:
1659: \bibitem{ttrev} Thiemann T 2004 \emph{Modern canonical quantum
1660: general relativity} (Cambridge University Press, Cambridge, at
1661: press) \textit{Preprint} \texttt{gr-qc/0110034}
1662:
1663: \bibitem{mb3} Bojowald M 2001 Absence of singularity in loop
1664: quantum cosmology \textit{Phys. Rev. Lett.} \textbf{86} 5227--5230
1665:
1666: \bibitem{mb6} Bojowald M 2003 Homogeneous loop quantum
1667: cosmology \textit{Class.\ Quantum Grav.} {\bf 20} 2595--2615
1668:
1669: \bibitem{abl} Ashtekar A, Bojowald M and Lewandowski J 2003
1670: \textit{Mathematical structure of loop quantum cosmology} {Adv.
1671: Theor. Math. Phys.} \textbf{7} 233--268
1672:
1673: \bibitem{aps} Ashtekar A, Pawlowski T and Singh P 2005
1674: Quantum nature of the big bang: An analytical and numerical study
1675: \textit{Preprint} in preparation
1676:
1677: \bibitem{md} Dafermos M 2003 Black hole formation from a complete
1678: regular past \textit{Preprint} \texttt{gr-qc/0310040}
1679:
1680: \bibitem{hm} Horowitz G T and Myers R 1995 The value of
1681: singularities \textit{Gen. Rel. Grav.} \textbf{27} 915-919
1682:
1683: \bibitem{group} Marolf D 1995 Refined algebraic quantization:
1684: Systems with a single constraint \textit{Preprint}
1685: \texttt{gr-qc/9508015};
1686: Giulini N and Marolf D 1999 On the Generality of refined
1687: algebraic quantization \textit{Class. Quant. Grav.} \textbf{16}
1688: 2479--2488; A uniqueness theorem for constraint quantization
1689: \textit{Class. Quant. Grav.} \textbf{16} 2489--2505
1690:
1691: \bibitem{ab2} Ashtekar A and Bojowald M 2005 {Black hole
1692: evaporation: A paradigm} \textit{Class. Quant. Grav.} \textbf{22}
1693: 3349--3362
1694:
1695: \bibitem{s'thw} Stevens C R, 't Hooft G and Whiting B F 1994 Black
1696: hole evaporation without information loss, \textit{Class. Quantum.
1697: Grav.} \textbf{11} 621--648
1698:
1699: \bibitem{sh3} Hayward S 2005 The disinformation problem for black
1700: holes (conference version), \texttt{gr-qc/0504038}
1701:
1702: \bibitem{m}
1703: Modesto L 2004 The Kantowski-Sachs space-time in loop quantum
1704: gravity \textit{Preprint} \texttt{gr-qc/0411032}
1705:
1706: \bibitem{kn}
1707: Kobayashi S and Nomizu K 1963 {\em Foundations of differential
1708: geometry} (John Wiley, New York)
1709:
1710: \bibitem{b}
1711: Brodbeck O 1996 On symmetric gauge fields for arbitrary gauge and
1712: symmetry groups {\em Helv.\ Phys.\ Acta} {\bf 69} 321--324
1713:
1714: \bibitem{bk}
1715: Bojowald M and Kastrup H A 2000 Symmetry reduction for quantized
1716: diffeomorphism invariant theories of connections {\em Class.\
1717: Quantum Grav.} {\bf 17} 3009--3043
1718:
1719: \bibitem{as} Ashtekar A and Samuel J 1991 Bianchi cosmologies:
1720: The role of spatial topology \textit{Class. Quant.Grav.} \textbf{
1721: 8} 2191--2215
1722:
1723: \bibitem{kt}
1724: Kastrup H A and Thiemann T 1994 Spherically symmetric gravity as a
1725: completely integrable system {\em Nucl.\ Phys.\ B} {\bf 425}
1726: 665--686
1727:
1728: \bibitem{k}
1729: Kucha\v{r} K V 1994 Geometrodynamics of Schwarzschild black holes
1730: {\em Phys.\ Rev.\ D} {\bf 50} 3961--3981
1731:
1732: \bibitem{tt} Thiemann T 1996 {Anomaly-free formulation of
1733: non-perturbative, four-dimensional Lorentzian quantum gravity}
1734: \textit{Phys. Lett.} \textbf{B380} 257--264; 1998 {Quantum spin
1735: dynamics (QSD)} \textit{Class. Quant. Grav.} \textbf{15}
1736: 839--873; 1998 {QSD V : Quantum gravity as the natural regulator
1737: of matter quantum field theories} \textit{Class. Quant. Grav.}
1738: \textbf{15} 1281--1314
1739:
1740: \bibitem{afw} Ashtekar A, Fairhurst S and Willis J 2003
1741: Quantum gravity, shadow states, and quantum mechanics
1742: \textit{Class. Quantum Grav.} \textbf{20} 1031--1062
1743:
1744: \bibitem{jw} Willis J 2004 On the low-energy ramifications and
1745: a mathematical extension of loop quantum gravity \textit{Ph.D.
1746: Dissertation, Penn State}
1747:
1748: \bibitem{mb1} Bojowald M 2001 {Inverse scale factor in isotropic
1749: quantum geometry} \textit{Phys.\ Rev.} \textbf{D64}
1750: 084018
1751:
1752: \bibitem{mb10} Bojowald M 2005 Degenerate Configurations,
1753: Singularities and the Non-Abelian Nature of Loop Quantum Gravity
1754: {\em Preprint} \texttt{gr-qc/0508118}
1755:
1756: \bibitem{bt} Brunnemann J and Thiemann T 2005 Unboundedness of Triad-Like
1757: Operators in Loop Quantum Gravity {\em Preprint} \texttt{gr-qc/0505033}
1758:
1759: \bibitem{bdv} Bojowald M, Date G and Vandersloot K 2004
1760: {Homogeneous loop quantum cosmology: The role of the spin
1761: connection} \textit{Class. Quantum Grav.} \textbf{21}
1762: 1253--1278
1763:
1764: \bibitem{bd} Bojowald M and Date G 2004 {Consistency conditions
1765: for fundamentally discrete theories} \textit{Class. Quantum
1766: Grav.} \textbf{21} (2004) 121--143.
1767:
1768: \bibitem{mb7} Bojowald M 2002 {Isotropic loop quantum cosmology}
1769: \textit{Class. Quantum Grav.} {\bf 19} 2717--2741.
1770:
1771: \bibitem{abw} Ashtekar A, Bojowald M, Willis J 2005 Modifications
1772: to Friedmann's equation from quantum geometry {\em Preprint} in preparation
1773:
1774: \bibitem{mb8} Bojowald M 2001 The semiclassical limit of loop
1775: quantum cosmology {\em Class.\ Quantum Grav.} {\bf 18} L109--L116
1776:
1777: \bibitem{mb4} Bojowald M 2001 Dynamical initial conditions in quantum
1778: cosmology {\em Phys.\ Rev.\ Lett.} {\bf 87} 121301
1779:
1780: \bibitem{ck}
1781: Cartin D and Khanna G 2005
1782: Absence of pre-classical solutions in Bianchi I loop quantum cosmology
1783: {\em Phys.\ Rev.\ Lett.} {\bf 94} 111302
1784:
1785: \bibitem{gd}
1786: Date G 2005
1787: Pre-classical solutions of the vacuum Bianchi I loop quantum cosmology
1788: {\em Preprint} \texttt{gr-qc/0505030}
1789:
1790: \bibitem{aps2} Ashtekar A, Pawlowski T and Singh P 2005
1791: Closed cosmology and quantum geometry {\em Preprint} in preparation
1792:
1793: \bibitem{mbss} Bojowald M 2004 Spherically symmetric quantum geometry:
1794: states and basic operators \textit{Class.\ Quantum Grav.}
1795: \textbf{21} 3733--3753;
1796: Bojowald M and Swiderski R 2004 The volume operator in spherically
1797: symmetric quantum geometry \textit{Class.\ Quantum Grav.}
1798: \textbf{21} 4881--4900;
1799: Bojowald M and Swiderski R 2005 Spherically symmetric quantum geometry:
1800: Hamiltonian constraint {\em Preprint} in preparation
1801:
1802: \bibitem{hw}
1803: Husain V and Winkler O 2004 Quantum resolution of black hole
1804: singularities \textit{Preprint} \texttt{gr-qc/0410125}
1805:
1806: \bibitem{mb9}
1807: Bojowald M 2005 Non-singular black holes and degrees of freedom in
1808: quantum gravity {\em Phys.\ Rev.\ Lett.} {\bf 95} 061301
1809:
1810: \end{thebibliography}
1811:
1812: \end{document}
1813:
1814:
1815:
1816:
1817: non-symmetric constraint
1818:
1819: \begin{eqnarray}
1820: && (V_{\mu+3\delta,\nu+2\delta}-V_{\mu+\delta,\nu+2\delta}) \psi_{\mu+2\delta,\nu+2\delta}
1821: -(V_{\mu-\delta,\nu+2\delta}-V_{\mu-3\delta,\nu+2\delta})
1822: \psi_{\mu-2\delta,\nu+2\delta}\nonumber\\
1823: && +\case{1}{2}(V_{\mu+4\delta,\nu+\delta}-V_{\mu+4\delta,\nu-\delta})
1824: \psi_{\mu+4\delta,\nu}
1825: -(1+\gamma^2\delta^2) (V_{\mu,\nu+\delta}-V_{\mu,\nu-\delta})
1826: \psi_{\mu,\nu}\nonumber\\
1827: && \qquad+\case{1}{2}(V_{\mu-4\delta,\nu+\delta}-V_{\mu-4\delta,\nu-\delta})
1828: \psi_{\mu-4\delta,\nu}\nonumber\\
1829: && -(V_{\mu+3\delta,\nu-2\delta}-V_{\mu+\delta,\nu-2\delta})
1830: \psi_{\mu+2\delta,\nu-2\delta}
1831: +(V_{\mu-\delta,\nu-2\delta}-V_{\mu-3\delta,\nu-2\delta})
1832: \psi_{\mu-2\delta,\nu-2\delta}\nonumber\\
1833: &=& 2\delta\sqrt{|\nu+2\delta|} (\psi_{\mu+2\delta,\nu+2\delta}-
1834: \psi_{\mu-2\delta,\nu+2\delta})\nonumber\\
1835: && +\case{1}{2} (\sqrt{|\nu+\delta|}-\sqrt{|\nu-\delta|})
1836: \left((\mu+4\delta)\psi_{\mu+4\delta,\nu}-
1837: 2(1+\gamma^2\delta^2)\mu\psi_{\mu,\nu}+
1838: (\mu-4\delta)\psi_{\mu-4\delta,\nu}\right)\nonumber\\
1839: &&-2\delta\sqrt{|\nu-2\delta|}(\psi_{\mu+2\delta,\nu-2\delta}-\psi_{\mu-2\delta,\nu-2\delta})\nonumber\\
1840: &=& 0\,. \label{diffeq2}
1841: \end{eqnarray}
1842:
1843:
1844:
1845:
1846:
1847: \bibitem{mb2} M. Bojowald, \textit{Loop quantum cosmology IV: Discrete
1848: time evolution} {Class. Quant. Grav.}, {\bf 18} (2001),
1849: 1071--1088.
1850:
1851: \bibitem{mb3} M. Bojowald, \textit{Absence of singularity in loop quantum
1852: cosmology}, {Phys. Rev. Lett.}, \textbf{86} (2001), 5227--5230.
1853:
1854: \bibitem{mb4} M. Bojowald, \textit{Dynamical initial conditions in quantum
1855: cosmology}, {Phys. Rev. Lett.}, {\bf 87} (2001), 121301.
1856:
1857: \bibitem{mb4a} M. Bojowald, \textit{Inflation from quantum geometry},
1858: {Phys. Rev. Lett.}, {\bf 89} (2002), 261301.
1859:
1860: \bibitem{mb5} M. Bojowald, \textit{Quantization ambiguities in
1861: isotropic quantum geometry}, {Class. Quantum Grav.}, {\bf
1862: 19} (2002), 5113--5130.
1863:
1864: \bibitem{ai} A. Ashtekar and C. J. Isham, \textit{Representation
1865: of the holonomy algebras of gravity and non-Abelian gauge
1866: theories}, {Class. Quant. Grav.}, \textbf{9} (1992), 1433--1467.
1867:
1868: \bibitem{al2} A. Ashtekar and J. Lewandowski, \textit{Representation
1869: theory of analytic holonomy algebras}, in `Knots and Quantum
1870: Gravity', edired by J.C. Baez, Oxford U.\ Press, Oxford 1994.
1871:
1872: \bibitem{jb1} J.C. Baez, J.C. \textit{Generalized measures in gauge
1873: theory}, {Lett. Math. Phys.}, \textbf{31} (1994), 213-223.
1874:
1875: \bibitem{mm} D. Marolf and J. Mour\~ao, On the support of the
1876: Ashtekar-Lewandowski measure, \textit{Commun. Math. Phys.},
1877: \textbf{170} (1995), 583-606.
1878:
1879: \bibitem{al3} A. Ashtekar and J. Lewandowski, \textit{
1880: Projective techniques and functional integration}, {Jour. Math.
1881: Phys.}, \textbf{36} (1995), 2170-2191.
1882:
1883: \bibitem{al4} A. Ashtekar and J. Lewandowski, \textit{Differential
1884: geometry on the space of connections using projective techniques},
1885: {Jour. Geo. \& \ Phys.}, \textbf{17} (1995), 191--230.
1886:
1887: \bibitem{al8} A. Ashtekar and J. Lewandowski,\textit{Background
1888: independent quantum gravity: A status report} (pre-print, 2003)
1889:
1890: \bibitem{aa1} A. Ashtekar, \textit{New variables for classical and
1891: quantum gravity}, {Phys. Rev. Lett.}, \textbf{57} (1986), 2244--2247;\\
1892: \textit{New Hamiltonian formulation of general relativity}, {Phys.
1893: Rev.}, \textbf{D36} (1987), 1587--1602.
1894:
1895: \bibitem{al5} A. Ashtekar and J. Lewandowski, \textit{Quantum theory of
1896: geometry I: Area operators}, {Class. Quant. Grav.}, \textbf{14}
1897: (1997), A55--A81.
1898:
1899: \bibitem{almmt} A. Ashtekar, J. Lewandowski, D. Marolf, J. Mour\~ao,
1900: and T. Thiemann, \textit{Quantization of diffeomorphism invariant
1901: theories of connections with local degrees of freedom}, {Jour.
1902: Math. Phys.}, \textbf{36} (1995), 6456--6493.
1903:
1904: \bibitem{hs} H. Sahlmann, \textit{Some comments on the representation
1905: theory of the algebra underlying loop quantum gravity},
1906: \texttt{gr-qc/0207111}.
1907: \bibitem{lo} J. Lewandowski and A. Okolow, \textit{Diffeomorphism
1908: covariant representations of the holonomy-flux star-algebra},
1909: \texttt{gr-qc/0302059}.
1910: \bibitem{st} H. Sahlmann and T. Thiemann, \textit{On the
1911: superselection theory of the Weyl algebra for diffeomorphism
1912: invariant quantum gauge theories}, \texttt{gr-qc/0302090}.
1913:
1914: \bibitem{afw} A. Ashtekar, S. Fairhurst and J. Willis,
1915: \textit{Quantum gravity, shadow states, and quantum mechanics},
1916: {Class. Quantum Grav.}, \textbf{20} (2003), 1031-1062.
1917:
1918: \bibitem{rs} C. Rovelli and L. Smolin, \textit{Spin networks and
1919: quantum gravity}, {Phys. Rev.}, \textbf{D52} (1995), 5743--5759.
1920:
1921: \bibitem{jb2} J. C. Baez, \textit{Spin networks in non-perturbative
1922: quantum gravity}, in `The Interface of Knots and Physics', edited
1923: by Kauffman, L. (American Mathematical Society, Providence, 1996)
1924: pp. 167--203;\\
1925: \textit{Spin networks in gauge theory}, {Adv. Math.} {\bf 117}
1926: (1996), 253--272.
1927:
1928: \bibitem{mb7} M. Bojowald, \textit{Isotropic loop quantum cosmology},
1929: {Class. Quantum Grav.}, {\bf 19} (2002), 2717--2741.
1930:
1931: \bibitem{abl} A. Ashtekar, M. Bojowald and J. Lewandowski,
1932: \textit{Mathematical Structure of Loop Quantum Cosmology},
1933: {Adv. Theor. Math. Phys.}, \textbf{7} (2003) 233--268.
1934:
1935: \bibitem{abck} A. Ashtekar, J. Baez, A. Corichi and K. Krasnov,
1936: \textit{Quantum geometry and black hole entropy}, {Phys. Rev.
1937: Lett.}, \textbf{80} (1998), 904--907.\\
1938: A. Ashtekar, J. Baez and K. Krasnov, \textit{Quantum geometry of
1939: isolated horizons and black hole entropy}, {Adv. Theo. Math.
1940: Phys.}, \textbf{4} (2000), 1--95.
1941:
1942: \bibitem{bdv} M. Bojowald, G. Date and K. Vandersloot,
1943: \textit{Homogeneous loop quantum cosmology: The role of the spin
1944: connection}, {Class. Quantum Grav.}, \textbf{21} (2004) 1253--1278.
1945:
1946: \bibitem{bd} M. Bojowald and G. Date, \textit{Consistency Conditions
1947: for Fundamentally Discrete Theories}, {Class. Quantum Grav.},
1948: \textbf{21} (2004) 121--143.
1949:
1950: \bibitem{mb8} M. Bojowald, \textit{The semiclassical limit of loop
1951: quantum cosmology}, {Class. Quantum Grav.}, \textbf{18} (2002),
1952: L109--L116.
1953:
1954: \bibitem{gr} M. Gaul and C. Rovelli, \textit{A generalized Hamiltonian
1955: constraint operator in loop quantum gravity and its simplest
1956: Euclidean Matrix Elements}, {Class. Quantum Grav.}, \textbf{18}
1957: (2001), 1593--1624.
1958:
1959: \bibitem{area} C. Rovelli and L. Smolin, \textit{Discreteness of area
1960: and volume in quantum gravity}, {Nucl. Phys.} \textbf{B442}
1961: (1995), 593--622; Erratum: {Nucl. Phys.} \textbf{B456} (1995)
1962: 753\\
1963: S. Frittelli, L. Lehner and C. Rovelli (1996) \textit{The complete
1964: spectrum of the area from recoupling theory in loop quantum
1965: gravity}, {Class. Quant. Grav.} \textbf{13}, 2921--2932
1966:
1967: \bibitem{dm} Marolf, D. \textit{Refined algebraic quantization:
1968: Systems with a single constraint}, \texttt{gr-qc/9508015}.
1969:
1970: \bibitem{tt1} T. Thiemann, \textit{QSD III: Quantum constraint
1971: algebra and physical scalar product in quantum general
1972: relativity}, {Class. Quantum Grav.} \textbf{15} (1998),
1973: 1207--1247.
1974:
1975: \bibitem{bk} M. Bojowald, and H. ~A. Kastrup, \textit{Symmetry
1976: reduction for quantized diffeomorphism invariant theories of
1977: connections}, {Class.\ Quantum Grav.}, {\bf 17} (2000),
1978: 3009--3043.
1979:
1980: \bibitem{mb9} M. Bojowald, \textit{Loop quantum cosmology:
1981: I. Kinematics}, \textit{Class.\ Quantum Grav.}, {\bf 17} (2001),
1982: 1489--1508.
1983:
1984: %\bibitem{tt2} T. Thiemann, \textit{Kinematical Hilbert spaces for
1985: %fermionic and Higgs quantum field theories}, {Class.\ Quantum
1986: %Grav.}, {\bf 15} (1998), 1487--1512.
1987:
1988: \bibitem{als} A. Ashtekar, J. Lewandowski and H. Sahlmann,
1989: \textit{Polymer and Fock representations for a Scalar field},
1990: {Class.\ Quantum Grav.} {\bf 20} (2003), L1--11.
1991:
1992: \bibitem{gp} R. Gambini and J. Pullin, \textit{
1993: Discrete quantum gravity: applications to cosmology},
1994: \texttt{gr-qc/0112033}.
1995:
1996: \bibitem{ae} A. Erd\'elyi, \textit{Asymptotic
1997: expansions}, Chapter II (Dover, New York, 1956).
1998:
1999: \bibitem{hm}
2000: G.~T. Horowitz and J. Maldacena, \textit{The black hole final state},
2001: JHEP {\bf 0402} (2004), 008.
2002:
2003: \end{thebibliography}
2004:
2005:
2006: \end{document}
2007:
2008:
2009:
2010: \vspace{1cm}
2011:
2012: OLD:
2013:
2014: scalars $\phi_1=a\Lambda_1+b\Lambda_2$,
2015: $\phi_2=-b\Lambda_2+a\Lambda_1$, $\phi_3=c\Lambda_3$ yield
2016: configuration space given by point holonomies:
2017: \begin{equation}
2018: {\cal U}=\{(\exp(A\Lambda_1),\exp(A\Lambda_2),\exp(c\Lambda_3))\}
2019: \end{equation}
2020:
2021: gauge invariant states are functions $f(A,c)$ even in $A$
2022:
2023: measure (from triple product of Haar measure):
2024: \begin{equation}
2025: \int f(A,c)\sin^2\case{A}{2} \sin^2\case{c}{2}\md A\md c
2026: \end{equation}
2027:
2028: orthonormal basis:
2029: \begin{equation}
2030: \chi_{j_1,j_2}=\frac{\sin(j_1+\case{1}{2})A}{\sin\case{A}{2}}
2031: \frac{\sin(j_2+\case{1}{2})c}{\sin\case{c}{2}} \quad,\quad
2032: \zeta_{j_1,j_2}=\frac{\sin(j_1+\case{1}{2})A}{\sin\case{A}{2}}
2033: \frac{\cos(j_2+\case{1}{2})c}{\sin\case{c}{2}}
2034: \end{equation}
2035:
2036: or
2037:
2038: \begin{equation}
2039: |n,m\rangle=\frac{\exp(\case{1}{2}imA)\exp(\case{1}{2}inc)}{
2040: 2\sin\case{A}{2}\sin\case{c}{2}}
2041: \end{equation}
2042:
2043: gauge invariant only $|n,m\rangle-|n,-m\rangle$: without restriction $m\geq1$
2044:
2045: \subsection*{Basic Operators}
2046:
2047: \begin{equation}
2048: \hat{p}_A=\frac{1}{4}\left(\widehat{\Lambda_1^i
2049: J_{1i}}+\left(\widehat{\Lambda_1^i J_{1i}}\right)^*\right)\quad,\quad
2050: \hat{p}_c=\frac{1}{2}\left(\widehat{\Lambda_3^i
2051: J_{3i}}+\left(\widehat{\Lambda_3^i J_{3i}}\right)^*\right)
2052: \end{equation}
2053:
2054: action can be found by acting on $\tr(h_1^m)\tr(h_3^n)$ and
2055: $\tr(h_1^m)\tr(\Lambda_3h_3^n)$ where $h_1=\exp(A\Lambda_1)$,
2056: $h_3=\exp(c\Lambda_3)$ and symmetrization:
2057:
2058: \begin{equation}
2059: \hat{p}_A=-\case{1}{2}i(\case{\md}{\md
2060: A}+\case{1}{2}\cot\case{A}{2})\quad,\quad \hat{p}_c= -i
2061: (\case{\md}{\md c}+\case{1}{2}\cot\case{c}{2})
2062: \end{equation}
2063:
2064: \begin{equation}
2065: \hat{p}_A|n,m\rangle=\case{1}{4}m|n,m\rangle\quad,\quad
2066: \hat{p}_c|n,m\rangle=\case{1}{2}n|n,m\rangle
2067: \end{equation}
2068:
2069: volume operator: $\hat{V}=\sqrt{\Lambda_3^i
2070: J_{3i}J_1^iJ_{1i}}+(\ldots)^*$
2071:
2072: \begin{equation}
2073: \hat{V}|n,m\rangle=\case{1}{4}\sqrt{m^2-1}\sqrt{\case{1}{2}|n|}=:
2074: V_{j(n),j(m)}|n,m\rangle
2075: \end{equation}
2076:
2077: with $j(k):=(|k|-1)/2$
2078:
2079: (point) holonomies lead to
2080:
2081: \begin{equation}
2082: \cos\case{A}{2}|n,m\rangle=\case{1}{2}(|n,m+1\rangle+|n,m-1\rangle)
2083: \quad,\quad
2084: \sin\case{A}{2}|n,m\rangle=\case{1}{2i}(|n,m+1\rangle-|n,m-1\rangle)
2085: \end{equation}
2086:
2087: and analogously for $\cos\case{c}{2}$, $\sin\case{c}{2}$
2088:
2089: \subsection*{Inverse Triad Components}
2090:
2091: \begin{equation}
2092: m_{IJ}:=q_{IJ}/\sqrt{\det q}=\diag\left(\frac{\sqrt{|p_c|}}{|p_A|},
2093: \frac{\sqrt{|p_c|}}{|p_A|}, \frac{|p_A|}{|p_c|^{\frac{3}{2}}}\right)
2094: \end{equation}
2095:
2096: quantization (as in isotropic models) leads to eigenvalues
2097:
2098: \begin{equation}
2099: m_{11;n,m}=16\left(\sqrt{V_{j(n),j(m+1)}}-
2100: \sqrt{V_{j(n),j(m-1)}}\right)^2>0\quad (\mbox{due to }m\geq1)
2101: \end{equation}
2102:
2103: \begin{equation}
2104: m_{33;n,m}=16\left(\sqrt{V_{j(n+1),j(m)}}-
2105: \sqrt{V_{j(n+1),j(m)}}\right)^2 \mbox{ vanishes for }n=0
2106: \end{equation}
2107:
2108: large $m,|n|$:
2109: $m_{11;n,m}\sim\case{4}{m}\sqrt{|n|/2}\sim\frac{\sqrt{|p_c|}}{p_A}$ and
2110: $m_{33;n,m}\sim\case{m}{4}(|n|/2)^{-\frac{3}{2}}\sim
2111: \frac{p_A}{|p_c|^{\frac{3}{2}}}$
2112:
2113: quantized co-triad components with eigenvalues:
2114:
2115: \begin{equation}
2116: \hat{a}_A:=\sqrt{|\hat{p_c}|}\quad,\quad
2117: a_{A;n,m}=\sqrt{\case{1}{2}|n|}
2118: \end{equation}
2119:
2120: \begin{equation}
2121: \hat{a}_c:=\hat{m}_{33}\hat{p_c}\quad,\quad
2122: a_{c;n,m}= \sqrt{2} n\sqrt{m^2-1}
2123: \left(|n+1|^{\frac{1}{4}}-|n-1|^{\frac{1}{4}}\right)^2
2124: \end{equation}
2125:
2126: (use $\hat{m}_{33}$ since $\hat{p}_c$ not invertible)
2127:
2128: \subsection*{Comparison with Schwarzschild}
2129:
2130: $a_A=t$, $|a_c|=\sqrt{(2M-t)/t}$ (mass $M$)
2131:
2132: horizon at $t=2M\rightarrow a_A>0,a_c=0$: $n\not=0$, $m=1$
2133:
2134: singularity at $t=0\rightarrow a_A=0,a_c\to\infty$: $n=0$, $a_{c;0,m}$
2135: finite (as with inverse scale factor)
2136:
2137: \subsection*{Evolution}
2138:
2139: classical Euclidean constraint: $H^E=-\frac{\sgn
2140: p_c}{\sqrt{|p_c|(p_a^2+p_b^2)}} \left[(a^2+b^2-1)(p_a^2+p_b^2)+
2141: 2cp_c(ap_a+bp_b)\right]$
2142:
2143: Lorentzian: $H=-\gamma^{-2}\frac{\sgn
2144: p_c}{\sqrt{|p_c|(p_a^2+p_b^2)}} \left[(a^2+b^2+\gamma^2)(p_a^2+p_b^2)+
2145: 2cp_c(ap_a+bp_b)\right]$
2146:
2147: quantization of Euclidean part:
2148:
2149: \begin{eqnarray}
2150: \hat{H}^E &=& \frac{8i}{\gamma\kappa l_P^2}\left[
2151: 8\sin\case{A}{2}\cos\case{A}{2} \sin\case{c}{2}\cos\case{c}{2}
2152: (\sin\case{A}{2}\hat{V}\cos\case{A}{2}-
2153: \cos\case{A}{2}\hat{V}\sin\case{A}{2})\right.\\
2154: && - \left.(1-4\sin^2\case{A}{2}-2\sin^4\case{A}{2})
2155: (\sin\case{c}{2}\hat{V}\cos\case{c}{2}-
2156: \cos\case{c}{2}\hat{V}\sin\case{c}{2})\right]\nonumber
2157: \end{eqnarray}
2158:
2159: action in ``triad representation'' $s_{n,m}$ with $s_{n,-m}=-s_{n,m}$
2160: due to gauge invariance:
2161:
2162: \begin{eqnarray}
2163: (\hat{H}^Es)_{n,m} &=& \frac{2}{\gamma
2164: l_P^2}\left[\case{1}{2}\sqrt{|n+2|/2}
2165: \left(\case{1}{2}(\sqrt{(m+3)^2-1}-\sqrt{(m+1)^2-1})
2166: s_{n+2,m+2}\right.\right.\\
2167: &&\qquad -\left.
2168: \case{1}{2}(\sqrt{(m-1)^2-1}-\sqrt{(m-3)^2-1}) s_{n+2,m-2}\right)\nonumber\\
2169: &&- \case{1}{4}(\sqrt{|n+1|/2}-\sqrt{|n-1|/2})(\case{1}{4}\sqrt{(m+4)^2-1}
2170: s_{n,m+4}- 3\sqrt{(m+2)^2-1}s_{n,m+2}\nonumber\\
2171: &&\qquad+ \case{7}{2}\sqrt{m^2-1}s_{n,m}- 3\sqrt{(m-2)^2-1}s_{n,m-2}+
2172: \case{1}{4}\sqrt{(m-4)^2-1} s_{n,m-4})\nonumber\\
2173: &&- \case{1}{2}\sqrt{|n-2|/2}
2174: \left(\case{1}{2}(\sqrt{(m+3)^2-1}-\sqrt{(m+1)^2-1})
2175: s_{n-2,m+2}\right.\nonumber\\
2176: &&\qquad -
2177: \left.\left.\case{1}{2}(\sqrt{(m-1)^2-1}-\sqrt{(m-3)^2-1})
2178: s_{n-2,m-2}\right)\right]\nonumber
2179: \end{eqnarray}
2180:
2181: using $s_{n,-m}=-s_{n,m}$ if $1\leq m\leq 4$, e.g.\
2182:
2183: \begin{eqnarray}
2184: (\hat{H}^Es)_{n,1} &=& \frac{2}{\gamma
2185: l_P^2}\left[\case{1}{4}\sqrt{3}\sqrt{|n+2|/2}
2186: (\sqrt{5}s_{n+2,3}-s_{n+2,1})\right.\\
2187: &&- \case{1}{4}(\sqrt{|n+1|/2}-\sqrt{|n-1|/2})(\case{1}{2}\sqrt{6}
2188: s_{n,5}- \case{7}{2}\sqrt{2}s_{n,3})\nonumber\\
2189: &&- \left.\case{1}{4}\sqrt{3}\sqrt{|n-2|/2}
2190: (\sqrt{5}s_{n-2,3}-s_{n-2,1})\right]\nonumber
2191: \end{eqnarray}
2192:
2193: and
2194:
2195: \begin{eqnarray}
2196: (\hat{H}^Es)_{n,2} &=& \frac{2}{\gamma
2197: l_P^2}\left[\case{1}{2}(\sqrt{3}-1)\sqrt{|n+2|}
2198: s_{n+2,4}\right.\\
2199: &&- \case{1}{4}(\sqrt{|n+1|/2}-\sqrt{|n-1|/2})(\case{1}{4}\sqrt{35}
2200: s_{n,6}- 3\sqrt{15}s_{n,4}+ \case{13}{4}\sqrt{3} s_{n,2})\nonumber\\
2201: &&- \left.\case{1}{2}(\sqrt{3}-1)\sqrt{|n-2|}
2202: s_{n-2,4}\right]\nonumber
2203: \end{eqnarray}
2204:
2205: Evolution using $n$ as internal discrete time: from $n$ to $n+2$
2206: determine $\case{1}{2}(\sqrt{(m+3)^2-1}-\sqrt{(m+1)^2-1})
2207: s_{n+2,m+2}- \case{1}{2}(\sqrt{(m-1)^2-1}-\sqrt{(m-3)^2-1})
2208: s_{n+2,m-2}$ for all $m$ (modified formula for low $m$), solve
2209: recursively for $s_{n+2,m}$ by using boundary data $s_{n,1}$ and
2210: $s_{n,2}$ ``at the horizon'', evolution through $n=0$ (singularity)
2211: possible as in isotropic models
2212:
2213: Need boundary data at fixed $n_0$, $n_0+1$, $n_0+2$, $n_0+3$ {\em
2214: and\/} at $m=1,2$, analogous to Wheeler-DeWitt equation: contains
2215: $\frac{\partial}{\partial p_A}\frac{\partial}{\partial p_c}$, so needs
2216: boundary data at ``light cone'' in $(A,c)$-plane
2217:
2218: %%%%%%%%
2219:
2220:
2221: %The latter corresponds to the sign of the transversal
2222: %extrinsic curvature components of $K_{ab}$ since they are obtained
2223: %from the connection and the triad via
2224: %$K_{ab}=(\Gamma_a^i-A_a^i)e_b^i/\gamma$ (later we will see that
2225: %the connection components in this model are proportional to the
2226: %extrinsic curvature components themselves).
2227:
2228: In what follows we will fix the remaining gauge freedom:
2229:
2230: %In what follows we will choose the convention that the sign is carried
2231: %by the gauge invariant triad component even though it correponds to
2232: %the sign of extrinsic curvature. This has the advantage that the gauge
2233: %invariant connection component
2234: %\begin{equation}
2235: % A:=\sqrt{a^2+b^2}
2236: %\end{equation}
2237: %which by definition is always positive, does not depend on the triad
2238: %components. For the gauge invariant triad component we then have
2239: %\begin{equation}
2240: % p_A:=\frac{ap_a+bp_b}{\sqrt{a^2+b^2}}
2241: %\end{equation}
2242: %which can be positive or negative. It is important to keep in mind,
2243: %however, that the sign of $p_A$ does not influence the orientation of
2244: %the triad since that would be given by
2245: %$\sgn(p_cp_A^2)=\sgn(p_c)$. Instead, the sign information of $p_A$
2246: %corresponds to the sign of extrinsic curvature, as explained
2247: %above. (When viewing the system as a cosmological Kantowski--Sachs
2248: %model, putting the sign in the triad components has the additional
2249: %advantage that the recollapse can be studied in the triad
2250: %representation.)
2251:
2252: %ALTERNATIVELY: Thus, it is natural to put the sign
2253: %into the gauge invariant connection component by defining
2254: %\begin{equation}
2255: % A:=\sgn(ap_a+bp_b) \sqrt{a^2+b^2}\,.
2256: %\end{equation}
2257: %For the triad, on the other hand, we only need one sign factor in
2258: %order to indicate the orientation which is given by $\sgn(p_c)$. Thus,
2259: %we do not loose any information if be use the gauge invariant triad
2260: %component
2261: %\begin{equation}
2262: % p_A:=\sqrt{p_a^2+p_b^2}
2263: %\end{equation}
2264: %which by definition is always positive. In fact, the orientation would
2265: %be given by $\sgn(p_cp_A^2)$ which would not distinguish between
2266: %positive and negative $p_A$.
2267:
2268:
2269:
2270: \end{document}
2271: