gr-qc0505032/pa56.tex
1: \documentclass[12pt]{article}
2: \usepackage{amsmath,amsthm,latexsym,amssymb,amsfonts,epsfig,psfrag,mathrsfs,bbm}
3: 
4: 
5: \oddsidemargin -1cm
6: \evensidemargin -1cm
7: \topmargin -1.5cm
8: \textwidth 18cm  % 16
9: \textheight 24cm  % 24
10: %
11: %
12: \newcommand{\tiN}{\raisebox{-6.5pt}{$\displaystyle\stackrel{\displaystyle N}{\sim}$}}
13: \newcommand{\tiM}{\raisebox{-6.5pt}{$\displaystyle\stackrel{\displaystyle M}{\sim}$}}
14: 
15: \newcommand{\fixme}[1][]{{\flushright\bf{$\Longrightarrow$ #1 }}\marginpar{\textbf{FIXME}}}   
16: %mark as drafty written
17: \newcommand{\Neq}[1]{$\mathcal{N}=#1$}                             
18: \newcommand{\adss}{$AdS_5\times S^5$}
19: \newcommand{\lu}[1]{_{#1}\!\!}    
20: \newcommand{\lo}[1]{^{#1}\!\!} 
21: \newcommand{\rb}[1]{\raisebox{1.5ex}[0pt]{#1}}
22: \newcommand{\fcmt}[2]{\fbox{\mbox{\begin{minipage}[t]{#1cm} #2 \end{minipage}}}}
23: \newcommand{\cmt}[2]{\mbox{\begin{minipage}[t]{#1cm} #2 \end{minipage}}}
24: \newcommand{\mb}[1]{\mathbbm{#1}}  
25: \newcommand{\Muserfunction}[1]{A}               %Mathematica-Importe
26: \newcommand{\Mvariable}[1]{#1}
27: \newcommand{\ul}[1]{\underline{#1}}  
28: \makeatletter
29: \@addtoreset{equation}{section}
30: \makeatother
31: \renewcommand{\theequation}{\thesection.\arabic{equation}}
32: %
33: \pagestyle{plain}
34: %
35: % theorem counter
36: \setcounter{secnumdepth}{5}
37: \newtheorem{Theorem}{Theorem}[section]
38: \newtheorem{Definition}{Definition}[section]
39: \newtheorem{Lemma}{Lemma}[section]
40: \newtheorem{Corollary}{Corollary}[section]
41: \newtheorem{Proposition}{Proposition}[section]
42: %
43: %  technical abbreviations
44: \def\be{\begin{equation}}
45: \def\ee{\end{equation}}
46: \def\ba{\begin{eqnarray}}
47: \def\ea{\end{eqnarray}}
48: %
49: % abbreviations connected with spaces of connections
50: \def\a{{\cal A}}
51: \def\ab{\overline{\a}}
52: \def\ac{\a^\Cl}
53: \def\acb{\overline{\ac}}
54: \def\g{{\cal G}}
55: \def\gb{\overline{\g}}
56: \def\gc{\g^\Cl}
57: \def\gcb{\overline{\gc}}
58: \def\ag{\a/\g}
59: \def\agb{\overline{\ag}}
60: \def\agp{(\ag)_p}
61: \def\agpb{\overline{\agp}}
62: \def\agc{\ac/\gc}
63: \def\agcb{\overline{\agc}}
64: \def\agcp{(\agc)_p}
65: \def\agcpb{\overline{\agcp}}
66: \def\dprime{{\prime\prime}}
67: %
68: % naturals
69: \def\Nl{{\mathchoice
70: {\setbox0=\hbox{$\displaystyle\rm N$}\hbox{\hbox to0pt
71: {\kern0.4\wd0\vrule height0.9\ht0\hss}\box0}}
72: {\setbox0=\hbox{$\textstyle\rm N$}\hbox{\hbox to0pt
73: {\kern0.4\wd0\vrule height0.9\ht0\hss}\box0}}
74: {\setbox0=\hbox{$\scriptstyle\rm N$}\hbox{\hbox to0pt
75: {\kern0.4\wd0\vrule height0.9\ht0\hss}\box0}}
76: {\setbox0=\hbox{$\scriptscriptstyle\rm N$}\hbox{\hbox to0pt
77: {\kern0.4\wd0\vrule height0.9\ht0\hss}\box0}}}}
78: %
79: % integers
80: \def\Zl{{\mathchoice
81: {\setbox0=\hbox{$\displaystyle\rm Z$}\hbox{\hbox to0pt
82: {\kern0.4\wd0\vrule height0.9\ht0\hss}\box0}}
83: {\setbox0=\hbox{$\textstyle\rm Z$}\hbox{\hbox to0pt
84: {\kern0.4\wd0\vrule height0.9\ht0\hss}\box0}}
85: {\setbox0=\hbox{$\scriptstyle\rm Z$}\hbox{\hbox to0pt
86: {\kern0.4\wd0\vrule height0.9\ht0\hss}\box0}}
87: {\setbox0=\hbox{$\scriptscriptstyle\rm Z$}\hbox{\hbox to0pt
88: {\kern0.4\wd0\vrule height0.9\ht0\hss}\box0}}}}
89: %
90: % rationals
91: \def\Ql{{\mathchoice
92: {\setbox0=\hbox{$\displaystyle\rm Q$}\hbox{\hbox to0pt
93: {\kern0.4\wd0\vrule height0.9\ht0\hss}\box0}}
94: {\setbox0=\hbox{$\textstyle\rm Q$}\hbox{\hbox to0pt
95: {\kern0.4\wd0\vrule height0.9\ht0\hss}\box0}}
96: {\setbox0=\hbox{$\scriptstyle\rm Q$}\hbox{\hbox to0pt
97: {\kern0.4\wd0\vrule height0.9\ht0\hss}\box0}}
98: {\setbox0=\hbox{$\scriptscriptstyle\rm Q$}\hbox{\hbox to0pt
99: {\kern0.4\wd0\vrule height0.9\ht0\hss}\box0}}}}
100: %
101: % reals
102: \def\Rl{{\mathchoice
103: {\setbox0=\hbox{$\displaystyle\rm R$}\hbox{\hbox to0pt
104: {\kern0.4\wd0\vrule height0.9\ht0\hss}\box0}}
105: {\setbox0=\hbox{$\textstyle\rm R$}\hbox{\hbox to0pt
106: {\kern0.4\wd0\vrule height0.9\ht0\hss}\box0}}
107: {\setbox0=\hbox{$\scriptstyle\rm R$}\hbox{\hbox to0pt
108: {\kern0.4\wd0\vrule height0.9\ht0\hss}\box0}}
109: {\setbox0=\hbox{$\scriptscriptstyle\rm R$}\hbox{\hbox to0pt
110: {\kern0.4\wd0\vrule height0.9\ht0\hss}\box0}}}}
111: %
112: %complexes
113: \def\Cl{{\mathchoice
114: {\setbox0=\hbox{$\displaystyle\rm C$}\hbox{\hbox to0pt
115: {\kern0.4\wd0\vrule height0.9\ht0\hss}\box0}}
116: {\setbox0=\hbox{$\textstyle\rm C$}\hbox{\hbox to0pt
117: {\kern0.4\wd0\vrule height0.9\ht0\hss}\box0}}
118: {\setbox0=\hbox{$\scriptstyle\rm C$}\hbox{\hbox to0pt
119: {\kern0.4\wd0\vrule height0.9\ht0\hss}\box0}}
120: {\setbox0=\hbox{$\scriptscriptstyle\rm C$}\hbox{\hbox to0pt
121: {\kern0.4\wd0\vrule height0.9\ht0\hss}\box0}}}}
122: %
123: % quaternions
124: \def\Hl{{\mathchoice
125: {\setbox0=\hbox{$\displaystyle\rm H$}\hbox{\hbox to0pt
126: {\kern0.4\wd0\vrule height0.9\ht0\hss}\box0}}
127: {\setbox0=\hbox{$\textstyle\rm H$}\hbox{\hbox to0pt
128: {\kern0.4\wd0\vrule height0.9\ht0\hss}\box0}}
129: {\setbox0=\hbox{$\scriptstyle\rm H$}\hbox{\hbox to0pt
130: {\kern0.4\wd0\vrule height0.9\ht0\hss}\box0}}
131: {\setbox0=\hbox{$\scriptscriptstyle\rm H$}\hbox{\hbox to0pt
132: {\kern0.4\wd0\vrule height0.9\ht0\hss}\box0}}}}
133: %
134: % octonions
135: \def\Ol{{\mathchoice
136: {\setbox0=\hbox{$\displaystyle\rm O$}\hbox{\hbox to0pt
137: {\kern0.4\wd0\vrule height0.9\ht0\hss}\box0}}
138: {\setbox0=\hbox{$\textstyle\rm O$}\hbox{\hbox to0pt
139: {\kern0.4\wd0\vrule height0.9\ht0\hss}\box0}}
140: {\setbox0=\hbox{$\scriptstyle\rm O$}\hbox{\hbox to0pt
141: {\kern0.4\wd0\vrule height0.9\ht0\hss}\box0}}
142: {\setbox0=\hbox{$\scriptscriptstyle\rm O$}\hbox{\hbox to0pt
143: {\kern0.4\wd0\vrule height0.9\ht0\hss}\box0}}}}
144: %
145: 
146: 
147: 
148: \title{On (Cosmological) Singularity Avoidance in\\ Loop Quantum Gravity}
149: \author{
150: J. 
151: Brunnemann\thanks{jbrunnem@aei.mpg.de,jbrunnemann@perimeterinstitute.ca},
152: T. 
153: Thiemann\thanks{thiemann@aei.mpg.de,tthiemann@perimeterinstitute.ca}\\
154: \\
155: MPI f. Gravitationsphysik, Albert-Einstein-Institut, \\
156:            Am M\"uhlenberg 1, 14476 Potsdam, Germany\\
157: \\
158: and\\
159: \\
160: Perimeter Institute for Theoretical Physics,\\ 
161: 31 Caroline Street N, Waterloo, ON N2L 2Y5, Canada}
162: 
163: \date{{\small Preprint AEI-2005-098}}
164: 
165: 
166: 
167: \begin{document}
168: 
169: \maketitle
170: 
171: \begin{abstract}
172: Loop Quantum Cosmology (LQC), mainly due to Bojowald, is not the 
173: cosmological sector of Loop 
174: Quantum Gravity (LQG). Rather, LQC consists of a truncation of the phase 
175: space of classical General Relativity to spatially homogeneous situations 
176: which is then quantized by the methods of LQG. Thus, LQC is a quantum 
177: mechanical toy model (finite number of degrees of freedom) for LQG  
178: (a genuine QFT with an infinite number of degrees of freedom) which
179: provides important consistency checks. However, it is 
180: a non trivial question whether the predictions of LQC are robust after 
181: switching on the inhomogeneous fluctuations present in full LQG.
182: 
183: Two of the most spectacular findings of LQC are that 1. 
184: the inverse scale factor is bounded from above on zero volume 
185: eigenstates which hints at the avoidance of the local
186: curvature singularity and 2. that the Quantum Einstein Equations are
187: non -- singular which hints at the avoidance of the global  
188: initial singularity. 
189: This rests on 1. a key technique developed for LQG and 2. the fact 
190: that there are no inhomogeneous excitations. We display the result of a 
191: calculation for LQG which proves that 
192: the (analogon of the) inverse scale factor, while densely defined, is 
193: {\it not} bounded from above on zero volume eigenstates. Thus, in full 
194: LQG, if curvature singularity avoidance is
195: realized, then not in this simple way.
196: 
197: In fact, it turns out that the boundedness of the inverse scale factor is 
198: neither necessary nor sufficient for curvature singularity avoidance and 
199: that non -- singular evolution equations are neither necessary nor 
200: sufficient for initial singularity avoidance because none of these 
201: criteria are formulated in terms of observable quantities. 
202: After outlining what would be required, we present the results of a 
203: calculation for LQG which could be a first indication that our 
204: criteria at least for curvature singularity avoidance are satisfied in 
205: LQG.
206: \end{abstract}
207: 
208: 
209: \section{Introduction}
210: \label{s1}
211: 
212: 
213: Loop Quantum Gravity (LQG) is a candidate for a Quantum Field Theory (QFT)
214: in four dimensions which achieves to unify the principles of Quantum 
215: Theory (QT) and General Relativity (GR). This means that LQG implements 
216: the fundamental feature of GR, its background independence, in a quantum
217: setting, see e.g. \cite{1} for books and \cite{1a} for reviews. 
218: 
219: To see how radical that ambition is, recall that ordinary QFT 
220: relies on a background spacetime $(M,g_0)$. Here $M$ is a differential
221: manifold and $g_0$ is a prescribed background metric thereon. Ordinary QFT 
222: now axiomatically assumes causality with respect to $g_0$: Given two 
223: smeared field operators 
224: $\phi(f),\;\phi(f')$ such that the supports of the smearing functions 
225: $f,f'$ are spacelike separated (the non -- spacelike geodesics with 
226: respect to $g_0$ starting from supp$(f)$ never hit supp$(f')$ and vice 
227: versa) then the field operators must (anti)commute. Thus we see that 
228: the background metric $g_0$ pivotally finds its way into the very 
229: definition of the field algebra. Without $g_0$ we therefore do not even
230: know what a quantum field is (see however \cite{2} for an interesting 
231: reformulation of ordinary QFT which may get us rid of $g_0$). 
232: 
233: The background $g_0$, however, is incompatible with GR, in particular
234: Einstein's equations, which state that the metric $g$ is a dynamical field 
235: which cannot be prescribed but must be determined dynamically according
236: to the matter energy momentum density. Hence, ordinary QFT  
237: neglects backreaction effects from matter on geometry (which is 
238: appropriate in scattering situations of negible particle numbers). 
239: Worse, if $g$ becomes itself a quantum field, in particular if there 
240: is no background $g_0$ such that the fluctuations $g-g_0$ are small  
241: in an appropriate sense, then ordinary QFT fails to be a valid description 
242: of the physical processes\footnote{The alternative, to quantize the 
243: fluctuations themselves and to construct a graviton QFT on $g_0$ fails 
244: due to 1. the non -- renormalizability of GR (see however \cite{3} 
245: for ideas concerning a non -- trivial UV fix point) and 2. 
246: in violent situations 
247: there may simply be no $g_0$ at all such that the fluctuations are 
248: small.}. It is precisely in those situations that a 
249: full fledged quantum theory of both matter and geometry must take over 
250: and it is almost granted that these situations arise for instance in the 
251: early universe. 
252: 
253: Thus, the task of a quantum gravity theory must be to define a quantum 
254: field just on a differential manifold $M$ and not on a background 
255: spacetime $(M,g_0)$. This means that one has to throw many aspects of 
256: conventional QFT over board. LQG is trying to precisely achieve this goal. 
257: As one might expect, given that already classical GR is a rather difficult
258: classical field theory, this is a very hard enterprise, both conceptually 
259: and technically. Despite these difficulties, LQG has made steady progress 
260: over the past fifteen years and 
261: we are getting closer and closer 
262: to being able to answer the question 
263: whether LQG really is a QFT of GR in the sense that 1. classical GR is the 
264: classical limit of LQG and 2. when the fluctuations of the quantum metric 
265: are small, then the framework of QFT on $(M,g_0)$ is recovered. In other
266: words, if these questions could be answered affirmatively, then we would
267: have a viable quantum gravity theory in front of us and we could start 
268: drawing conclusions and physical predictions from it and hopefully compare 
269: those with experiments.
270: 
271: However, even before we (hopefully) complete the construction of full LQG, 
272: it is possible and mandatory to test the theory by models which are 
273: conceptually and technically easier to handle and which still capture 
274: enough of the features of the full theory. To give an example, the 
275: hydrogen atom can be analyzed just using quantum mechanics for the 
276: electrons orbiting a classical nucleus and one gets fantastically close 
277: to the experiment. Hence, this gives us an important first idea. However, 
278: the full problem of the hydrogen atom must certainly
279: take into account the full QFT of the standard model\footnote{Even 
280: today that has not yet been done fully, usually one only considers QED 
281: effects but not e.g. the QCD physics of the proton.} and only then one 
282: can determine the Feinstruktur of the hydrogen atom energy spectrum such 
283: as the Zitterbewegung of the electron due to the interaction with the 
284: electromagnetic field. 
285: 
286: It is remarkable and important that the predictions
287: of the simple quantum mechanical computation are confirmed by the full 
288: theory because a priori it could be that the full QED calculation 
289: completely changes the picture. Of course, in the case of the hydrogen 
290: atom one has access to experimental data and so this was not expected if 
291: QED has something to do with nature. However, in case of absence of data
292: there is no reason why one should trust a toy model calculation at all.
293: In particular, it is unclear why a quantum mechanical calculation
294: describing only a finite number of degrees of freedom will reproduce the 
295: QFT aspects of the full problem involving an infinite number of degrees of 
296: freedom. One hopes of course that the full problem will only yield 
297: corrections to the model, however, that is far from granted and must be 
298: proved rigorously.
299: 
300: As far as LQG is concerned, in recent years a new class of toy models 
301: have been constructed which were coined Loop Quantum Cosmology (LQC).
302: These developments are mainly due to Bojowald, see e.g. the 
303: beautiful review 
304: \cite{4}. These models consider the truncation of the infinite dimensional 
305: phase space of GR to spatially homogeneous situations which is finite 
306: dimensional. Such truncations have been considered before and are called 
307: minisuperspace models \cite{5}. Traditionally they were quantized using 
308: the Schr\"odinger representation of the canonical commutation relations.
309: What is new in Bojowald's work is that he employs a different 
310: representation of the canonical commutation relations. This might come as 
311: a surprise because the Stone von Neumann theorem gives us a uniqueness 
312: result concerning the representation theory of the Weyl algebra for 
313: finite dimensional phase spaces. The catch
314: is that the Stone von Neumann theorem assumes, among other things, (weak) 
315: continuity of the Weyl operators. Bojowald drops this assumption because 
316: in full LQG the unique \cite{5b} representation \cite{5a} is also 
317: discontinuous, hence this 
318: representation is a much better model for LQG than the usual Schr\"odinger
319: representation. This kind of representations were previously discussed by 
320: e.g. Thirring \cite{6} and while exotic are used e.g. in solid state 
321: physics \cite{7}. Moreover, using those one can circumvent negative norm 
322: states in Maxwell theory \cite{8} or string theory which then 
323: does not lead to restrictions on the spacetime dimension \cite{9}.
324: 
325: A second important ingredient in Bojowald's work is a key technique 
326: discovered for full LQG in \cite{10} which allows to define certain 
327: operators densely on the LQG Hilbert space although their classical 
328: counterpart is ill defined when the metric becomes degenerate. Prime 
329: examples for such classically singular situations are cosmological big 
330: bang singularities. This key technique was used in \cite{10} in order to 
331: give, for the 
332: first time, a mathematical meaning to the Quantum Einstein Equations in 
333: full LQG, mathematically speaking, the so -- called Wheeler -- DeWitt 
334: equations could be defined densely on the LQG Hilbert space. Moreover,
335: these equations can be solved explicitly by following a constructive 
336: algorithm \cite{10a}. This works for arbitrary matter coupling \cite{10b}
337: and was tested successfully in 2+1 gravity \cite{10c}. 
338: For a modern version 
339: of this framework see \cite{11}. A simpler setting in which this key 
340: technique was used is the construction of the length operator in full LQG
341: \cite{11a}. This operator is also well (i.e. densely) defined, it is an 
342: unbounded but positive self -- adjoint operator whose spectrum is 
343: entirely discrete. 
344: 
345: Of course, due to the complexity of the full theory, LQG cannot be solved 
346: completely. It is precisely the virtue of toy models such as LQC that 
347: their mathematics is comparatively simple so that one can focus 
348: immediately on  
349: the conceptual analysis. Bojowald therefore could go very far in the 
350: quantization of the cosmological models although some important aspects 
351: such as the physical inner product are still missing. The most 
352: spectacular finding is that in LQC the operator corresponding to the 
353: inverse scale factor(s)\footnote{One for the isotropic models, three for 
354: the diagonal models.} is (are) bounded from above \cite{12,13a}. 
355: This hints at the following: Recall that in classical 
356: GR one characterizes singularities by global and local criteria 
357: \cite{17a}. The global criterion is the causal geodesic incompleteness 
358: of a (inextendible) spacetime. Equivalently,
359: the Einstein equations on 
360: globally hyperbolic spactimes with manifolds diffeomorphic to 
361: $\Rl\times \sigma$ cannot be solved for all values of $t\in \Rl$.
362: The local criterion is defined as 
363: the divergence of polynomial scalars built out of the Riemann tensor
364: and is used to characterize a global singularity in more detail. The local 
365: criterion by itself is unsatisfactory because a 
366: spacetime can have a global singularity while all curvature scalars
367: are regular. Notice that both criteria are to be measured by an
368: obeserver and in this sense are to be formulated in terms of observables 
369: (measurable quantities).\\ 
370: In classical cosmology both types of singularities are present.
371: Since in cosmology curvature scalars are 
372: given by powers of the inverse scale factor, its boundedness hints at 
373: the absence of the local singularity in LQC in the sense that
374: with respect to any kinematical state the curvature expectation value 
375: can never diverge. Moreover, the 
376: truncated Quantum Einstein Equations (Wheeler DeWitt Equation)
377: do not become singular at the classical Big Bang and 
378: one therefore can go before the Big Bang. This hints 
379: at global singularity avoidance in LQC.
380: Further consequences of the 
381: boundedness of the inverse scale factor are that the quantum geometry 
382: behaves very non -- classically at the classical singularity implying 
383: effective modifications of the Friedmann equations which then drive {\it 
384: geometric inflation without the necessity of fine tuning the inflaton
385: potential} \cite{13}. These results related to the boundedness of the 
386: inverse scale factor in LQC are so promising and spectacular that it was 
387: extensively highlighted in the
388: physics literature and public press. In particular, the boundedness
389: was mentioned as crucial to establish the non singularity of the quantum 
390: evolution equations, see e.g. \cite{Bojowald1}.
391: 
392: In \cite{14} it was shown that it is precisely those two ingredients
393: imported from full LQG into LQC,
394: 1. the non -- standard representation of the canonical commutation 
395: relations and 2. the key technique \cite{10} to densely define classically 
396: degenerate operators, which are responsible for the absence of the 
397: local singularity in LQC and similar models inspired more by a 
398: metric than 
399: connection based approach. However, as outlined above, on general grounds
400: one must critically examine the stability of such model calculations when
401: one reintroduces the field theory degrees of freedom. This has been 
402: stressed before for geometrodynamics in \cite{14a} where one generically 
403: finds that the model even quantitatively has large deviations from the 
404: full theory or more complete theory. More recently also LQC has been 
405: examined critically   
406: \cite{14b} concerning the issue of the semiclassical limit but here the 
407: results seem to be inconclusive so far. 
408: 
409: The main purpose of this paper 
410: is to contribute to the stability analysis of LQC versus LQG.\\
411: \\
412: The paper is organized as follows:\\
413: \\ 
414: \\
415: In section two we summarize and compare the main features of LQG and LQC
416: for the unfamiliar reader. Experts can safely skip this section.\\
417: \\
418: In section three we report the results of a computation carried out in our 
419: companion paper \cite{16} within full LQG which 
420: is analogous to the one carried out in LQC. In LQC this calculation proves 
421: the boundedness of the inverse scale factor. 
422: More precisely, we ask, as in LQC, whether the
423: part of the energy momentum operator which probes the quantum metric is 
424: bounded from above on the kinematical Hilbert space of full LQG, at 
425: least at the classical singularity, that is, when the volume of space 
426: vanishes. This precisely mirrors the calculation of \cite{12}.
427: The result is negative: {\bf In full LQG the (analogon of) the inverse 
428: scale factor is unbounded from above even when the quantum volume 
429: vanishes.} 
430: This is true even when we compute its norm with respect to a 
431: state of zero volume which is homogeneous and isotropic on large 
432: scales. This proves that the boundedness of the inverse scale factor
433: in isotropic and homogeneous LQC does not extend to the full theory even
434: when restricting LQG to those states which one would use to
435: describe a maximally homogeneous and isotropic situation (modulo
436: fluctuations). 
437: 
438: It is true that in non -- isotropic, homogeneous
439: models the inverse scale factor is also unbounded on the full Hilbert
440: space. However, in those more general models it is still bounded on zero
441: volume eigenstates which again seems to indicate that the local 
442: singularity is evaded in LQC. Our result is stronger: No matter whether 
443: homogeneous
444: and/or isotropic, spherically symmetric, cylindrically symmetric etc., the
445: inverse scale factor is unbounded on zero volume
446: eigenstates in full LQG and on the full Hilbert space anyway. The reason 
447: for this
448: is that LQG, in contrast to LQC, admits inhomogeneous, microscopical
449: excitations which precisely account for the
450: unboundedness. We will see this explicitly when we construct the
451: zero volume eigenstates on which the inverse scale factor has
452: arbitrarily large norm. Notice that
453: meanwhile also inhomogeneous models such as spherically symmetric
454: or cylindrically symmetric ones have been quantized by LQG methods in which
455: the inverse scale factor apparently also stays bounded 
456: \cite{Bojowald}.
457: While these do have local, inhomogeneous degrees of freedom
458: (in the spherically symmetric case without matter, at least before
459: solving the constraints),
460: as our calculation reveals, the inhomogeneous excitations of
461: full LQG are more general than in those models and thus give unboundedness
462: even on states of LQG describing a sector appropriate for those 
463: more general models. \\
464: \\
465: However, this does not mean that LQG does not predict the absence of the
466: local initial singularity. Namely, in section four we remark that 
467: an operator may be unbounded but still it may be bounded when restricted 
468: to a subspace of the Hilbert space. Thus, the boundedness of the inverse 
469: scale factor is not a necessary criterion for local singularity avoidance. 
470: In our case we are interested in 
471: a sector which describes a collapsing universe which is homogeneous and 
472: isotropic at large scales. We perform a corresponding calculation
473: which may be taken as an indication that 
474: the expectation value of the analogon of the inverse scale factor in LQG, 
475: with respect to kinematical\footnote{Kinematical states, in contrast to 
476: physical states, do not solve the Quantum Einstein Equations. 
477: Non -- gauge invariant operators such as the inverse scale factor can 
478: only be probed on the kinematical Hilbert space.}, 
479: coherent states \cite{20} peaked on homogeneous and 
480: isotropic initail data, {\bf is bounded from above at the 
481: Big Bang}\footnote{Notice that coherent states peaked
482: on homogeneous and isotropic initial data still have inhomogeneous and 
483: anisotropic excitations, however, they are small.}. We did this for scalar 
484: matter only but qualitatively nothing changes in more general 
485: situations. Also our result is completely general: We derive a bound of 
486: the inverse scale factor on coherent states peaked on an arbitrary point 
487: in phase space. This formula can then be specialized also to inhomogeneous 
488: models and for the Bianchi I (Kasner) case it is easy to see that one gets 
489: boundedness as well.\\
490: \\
491: In section five we stress that the results of section four are promising 
492: but 
493: inconclusive: Unfortunately, the boundedness of the inverse 
494: scale factor is also not a sufficient criterion for 
495: local singularity avoidance. 
496: This has to do with the fact that both singularity criteria 
497: fundamentally have to be discussed at the level of physical 
498: observables and physical states. They must be be disussed 
499: separately because the quantum theory could avoid one of them but not the
500: other.\\ 
501: Concerning the local singularity, the inverse scale factor is not gauge 
502: invariant and therefore not an observable. It can be turned into an 
503: observable using the technique of partial observables \cite{15,15a} but 
504: then its spectrum on the physical Hilbert space can differ drastically 
505: from the kinematical spectrum \cite{20a}. Hence, reliable
506: statements about the local singularity can only be obtained on the 
507: physical Hilbert space.\\ 
508: As far as the global singularity is concerned, physical states 
509: automatically solve the quantum Einstein equations so there seems to be 
510: no sign of the global singularity at the level of physical states, by 
511: construction\footnote{Of course, the inverse scale factor must be densely 
512: defined to even define the Quantum Einstein Equations but that has been 
513: established for LQG in \cite{10,10b}.}. More precisely, in order to solve 
514: the Quantum Einstein Equations, one takes an arbitrary superposition 
515: of kinematical states and then obtains a recursion relation for the 
516: coefficients. These have been derived explicitly for full LQG in 
517: \cite{10a} and those of LQC \cite{12} are similar but simpler.
518: One of the labels of these coefficients can be identified with an 
519: unphysical time parameter. If one chooses that label in such a way 
520: that its range is 
521: (a discrete subset of) the full real axis, a solution to the Quantum 
522: Einstein Equations, that is, a physical state,
523: describes an entire quantum universe (a history) and involves the full 
524: range of the unphysical time parameter. {\it Thus there is no sign of the 
525: global singularity within a given solution even in the full theory}.
526: However, following \cite{12} one may argue that the existence
527: of a global singularity
528: could manifest itself in the fact that the recursion relations break down
529: at zero volume. Such a breakdown can manifest itself in the following 
530: two ways:\\
531: A.\\ 
532: It can lead
533: to a restriction or additional consistency condition on the
534: freedom in the choice of the initial
535: data (coefficients) that parameterize the solution space.
536: This makes the solution space smaller than it would be otherwise.
537: In particular, if the
538: number of additional conditions is larger than the free initial parameters
539: of the solution, the solution space would be empty.\\
540: B.\\
541: It can happen that certain coefficients remain undetermined by the
542: recursion relation and lead to an indeterministic quantum evolution
543: (given initial coefficients, we cannot construct the full solution
544: without making additional choices). This makes the solution space larger
545: than it would be otherwise. Notice that this undeterministic quantum
546: evolution corresponds to an undeterministic classical evolution (initial
547: value problem). However, by itself such ambiguities are not physically
548: relevant because this is not an evolution of physical quantities. Even 
549: classically it is an evolution with respect to an unphysical time of 
550: unphysical quantities. In terms of physical (gauge invariant) quantities 
551: {\it there is no unphysical evolution at all by definition} because Dirac
552: observables Poisson commute with the Hamiltonian constraint. Hence, 
553: what we actually have to do in order to observe a possible breakdown of 
554: evolution is to
555: consider an evolution with respect to a physical Hamiltonian
556: and of physical observables.
557: We will discuss this at length in section five.\\
558: We now point
559: out that absence of either effect A or B is neither a sufficient nor a 
560: necessary criterion for global singularity avoidance.\\
561: To see that it is not necessary, notice that the presence of either effect
562: indeed affects the size of the set of formal solutions to
563: the evolution equations. However, even if such effects
564: exist, it may still be the case that the physical Hilbert space is large
565: enough in order to accomodate all semiclassical physical states which
566: describe all
567: classical spacetimes (at large scales). To see that it is not
568: sufficient, notice that not
569: all formal solutions to the Quantum Einstein Equations correspond to
570: physical states because they might not be normalizable with respect to
571: the physical inner product or they might have zero norm.
572: In both cases we must drop them from the set of physical states.
573: Hence, even if effects A or B are absent, it could turn out, in the
574: worst case, that most of the solutions that one finds do not lie
575: in the physical Hilbert space and then again one would be confronted with
576: a global singularity because there could be not even be a semiclassical 
577: sector
578: describing cosmology. Furthermore, as argued above, global singularity 
579: avoidance 
580: must be formulated in terms of physical observables which has 
581: not been done yet. Therefore, the breakdown 
582: of the recursion or the absence thereof is inconclusive for the presence 
583: or absence of global singularities 
584: unless one knows the physical Hilbert space and the corresponding Dirac 
585: observables.
586:  
587: We conclude: {\bf In order to decide on the presence or absence of
588: both the local and global singularity, detailed knowledge of the physical 
589: Hibert space is a necessary 
590: prerequisite.}. This Hilbert space is known to 
591: exist both in LQG and LQC, following for instance \cite{11}. However,
592: the corresponding physical inner product is rather difficult to 
593: construct explicitly,
594: for any of the current versions of the Hamiltonian constraint,
595: which is why a definitive conclusion on the singularity issue
596: is not available at the moment.
597: 
598: To improve on this, we first of all broadly outline which steps should be 
599: fundamentally performed 
600: in order to prove the absence of both the global and  
601: local Big Bang singularity in full LQG. We give both a precise and 
602: an approximate scheme. 
603: The ideal (precise) procedure involves the exact
604: knowledge of the Hilbert space of physical states and a sufficiently 
605: large set of Dirac observables (gauge invariant quantities) which, as we 
606: said, are not 
607: yet available neither in full LQG nor in LQC in 
608: a form explicit enough. The
609: approximate scheme stays within the kinematical Hilbert space over which 
610: one has excellent control. The calculation performed in section 
611: four can be viewed as a simplified version of the approximate scheme. 
612: The application of either scheme is left for future research and again 
613: LQC would be an ideal testing ground for this scheme.\\
614: \\
615: We conclude in section six summarizing the status of the absence of 
616: the initial singularity in full LQG. Notice that qualitatively all we have 
617: said is also applicable to other singularities such as those connected 
618: with black holes \cite{20d,20c}. In fact, most considerations within LQG 
619: concerning black holes are based on 
620: minisuperspace calculations which exploit the fact that the interior of a 
621: black hole can be mapped into a homogeneous Kantowski -- Sachs model
622: so that one can transfer, almost literally, the LQC results. However, our 
623: remarks prevail: No reliable conclusions can be drawn before one works 
624: with physical operators and physical states of the full theory. This has 
625: also been stressed in \cite{20c}.\\
626: \\
627: Notice that this paper is not meant as a criticism of LQC by 
628: itself, in the contrary, LQC has provided us with many fruitful new 
629: conceptual ideas and one would hope that its results hold also LQG, albeit
630: in a technically different realization.
631: However, we want to draw attention to the fact that LQC is far from being 
632: full LQG and that therefore caution should be applied when trying to 
633: extrapolate LQC results to LQG. This extends also to more general reduced
634: models such as the spherically symmetric or cylindrically symmetric ones.
635: 
636: 
637: 
638: 
639: \section{Relation between LQG and LQC}
640: \label{s0}
641: 
642: For the benefit of the reader unfamiliar with the basic structure of 
643: either LQG or LQC we here summarize elements of both frameworks and 
644: compare them. The presentation will be oversimplified and the 
645: mathematical details will be discared.
646: As we will see, LQC does not describe a stable subsector
647: of LQG, rather it is a truncation in which all but finitely many degrees 
648: of 
649: freedom are cancelled by hand. Experts can safely skip this section.
650: 
651: \subsection{Elements of LQG}
652: \label{s0.1}
653: 
654: The classical starting point of LQG is a Hamiltonian formulation of the 
655: Einstein Hilbert action, however, written in unusual variables. Instead
656: of using the three metric $q_{ab}$ on the spatial slices $\sigma$ of a 
657: foliation of the four manifold $M\cong \Rl\times \sigma$ and the extrinsic 
658: curvature $K_{ab}$ as the canonical variables as advocated by Arnowitt,
659: Deser and Misner, one uses variables that are more famililar from the 
660: canonical formulation of Yang Mills theories. They consist of an 
661: $SU(2)$ connection $A_a^j$ and an electric field $E^a_j$. Here 
662: $a,b,c,..=1,2,3$ are tenorial indices and $j,k,l,..=1,2,3$ are $su(2)$
663: indices. The origin of $SU(2)$ is that $SU(2)$ is the universal covering 
664: group of $SO(3)$ (necessary for coupling of spinorial matter) and $SO(3)$
665: appears naturally when we write the three metric in terms of cotriads
666: (or frame fields) 
667: $e_a^j$, that is, $q_{ab}=e_a^j e_b^k \delta_{jk}$. Indeed, the cotriads 
668: are determined by $q_{ab}$ only up to an $SO(3)$ rotation and this is 
669: how $SU(2)$ finds its way into this formulation as a gauge group. 
670: 
671: The relation between $(A,E)$ and $(q,K)$ is 
672: \be \label{0.1}
673: A_a^j=\Gamma_a^j+\beta K_{ab} e^b_j,\;\;E^a_j=\sqrt{\det(q)} e^a_j
674: \ee
675: Here $\Gamma$ is the spin connection (a certain function of $e_a^j$ and 
676: its first spatial derivatives), the triad $e^a_j$ is the inverse of 
677: $e_a^j$, i.e. 
678: $e^a_j e^j_b=\delta^a_b,\;\; e^a_j e^k_a=\delta^k_j$ and $\beta>0$ is a 
679: parameter called the Immirzi parameter. That $(A,E)$ form a canonical pair 
680: means that
681: \be \label{0.2}
682: \{E^a_j(x),A_b^k(y)\}=8\pi G \beta \;\delta^a_b\;\delta^k_j\;\delta(x,y)
683: \ee
684: all others vanishing. The phase space coordinatized by $(A,E)$ is subject 
685: to the usual spatial diffeomorphism constraint which in these variables 
686: can be written as
687: \be \label{0.3}
688: C_a=F_{ab}^j\; E^b_j \mbox{ where } F_{ab}^j=\partial_a A_b^j-\partial_b 
689: A_a^j+\epsilon_{jkl} A_a^k A_b^l
690: \ee
691: is the curvature of $A$, the Hamiltonian constraint
692: \be \label{0.4}
693: C=\frac{\epsilon_{jkl} [F_{ab}^j-(1+\beta^2) \epsilon_{jmn} K_a^m K_b^n]
694: E^a_k E^b_l}{\sqrt{|\det(E)|}}
695: \ee
696: where $K_a^j=A_a^j-\Gamma_a^j(E)$ and the additional Gauss constraint
697: \be \label{0.5}
698: C_j=\partial_a E^a_j+\epsilon_{jkl} A_a^k E^a_l
699: \ee
700: which gets us rid of the $SU(2)$ degrees of freedom. We see that GR can be 
701: cast into the form of a $SU(2)$ Yang Mills theory (which is also subject 
702: to $C_j=0$) with one important difference: The background dependent Yang 
703: Mills Hamiltonian density
704: \be \label{0.6}
705: H=\frac{q^0_{ab}}{\sqrt{\det(q^0)}}\;[E^a_j E^b_k+B^a_j B^b_k]\delta^{jk}
706: \ee
707: where $B^a_j=\epsilon^{abc} F^j_{bc}/2$ is the magnetic field and
708: $q^0$ is the spatial projection of the background metric $g^0$,
709: is replaced by the background independent constraints $C_a,C$. 
710: 
711: If there is matter present such as a scalar field $\phi$ then the 
712: constraints $C_a,\; C$ are augmented by additional terms
713: \be \label{0.7}
714: C^{scalar}_a=\pi \phi_{,a} \mbox{ and } 
715: C^{scalar}=\frac{1}{2}\frac{\pi^2+E^a_j E^b_j 
716: \phi_{,a} \phi_{,b}}{\sqrt{|\det(E)|}}+\sqrt{|\det(E)|} V(\phi)
717: \ee
718: where $\pi$ is the momentum conjugate to $\phi$. 
719:  
720: Canonical quantization now, very roughly, means to find a Hilbert space 
721: ${\cal H}_{Kin}$ on which the canonical Poisson brackets (\ref{0.2})
722: are implemented as canonical commutation relations and to impose the 
723: quantum constraints, e.g. $\hat{C}\psi=0$. The resulting space of 
724: solutions is the physical Hilbert space on which the (Dirac) observables,
725: that is, the gauge invariant functions $F$ on phase space satisfying
726: $\{C_a(x),F\}=\{C(x),F\}=\{C_j(x),F\}=0$ act as self adjoint operators
727: $\hat{F}$. This simple recipe has to be refined in QFT because of the 
728: singular short distance behaviour of the fields $A,E$. In order to deal 
729: with this problem one has to smear the fields. There are many ways to smear 
730: the fields, each of which gives rise to a different algebra 
731: $\mathfrak{A}$ of elementary
732: observables\footnote{They are called elementary because one can write 
733: every function on phase space in terms of (limits of) them. They are 
734: not physically observable because they are not gauge invariant.}. Guided 
735: by experience with canonical lattice gauge field theory in LQG one 
736: considers the following objects:\\ 
737: 1. For each curve $e$ in $\sigma$
738: the holonomy of the connection $A$ along $e$, that is,
739: \be \label{0.8}
740: A(e):={\cal P}\;\exp(\int_e A):=1_2+\sum_{n=1}^\infty
741: \int_0^1\; dt_1\; \int_{t_1}^1\;dt_2\; ..\;\int_{t_{n-1}}^1\; dt_n\;
742: A(t_1)..A(t_n) 
743: \ee
744: where $e:\;[0,1] \to e; t \mapsto e(t)$ is a parameterization of the 
745: curve and $A(t):=A_a^j(e(t)) \dot{e}^a(t) \tau_j$ with 
746: $\tau_j=-i\sigma_j/2$ a basis of $su(2)$ ($\sigma_j$ are the Pauli 
747: matrices).\\
748: 2. For each surface $S$ and each $su(2)$ valued function $x\mapsto f^j(x)$
749: the electric flux of $E$ through $S$
750: \be \label{0.9}
751: E_f(S)=\int_s\; d^2u\; f^j(X_S(u))\; \epsilon_{abc} E^a_j(X_S(u))\;
752: X^b_{S,u^1}(u)\; X^c_{S,u^2}(u)
753: \ee
754: where $X_S:\;s\subset \Rl^2\to S;\;u=(u^1,u^2)\mapsto X_S(u)$ is a 
755: parameterization of 
756: $S$. \\
757: Using (\ref{0.2}) one can show that the Poisson brackets\footnote{More 
758: precisely the corresponding Hamiltonian vector fields.} of the holonomies 
759: and fluxes close among each other so that they define a well defined 
760: algebra $\mathfrak{A}$.  
761: 
762: The kinematical Hilbert space of LQG can now be described very easily as 
763: follows:\\
764: A union of curves $e$ forms a graph $\gamma$. By subdividing the curves we 
765: may assume that the curves are mutually disjoint except for the endpoints.
766: These curves are then called edges of $\gamma$ which intersect in the 
767: vertices. We will denote by $E(\gamma),\;V(\gamma)$ the set of 
768: edges and vertices of a graph respectively. The wave functions are then 
769: of the form 
770: \be \label{0.10}
771: \psi(A)=\psi_\gamma(A(e_1),..,A(e_N))
772: \ee
773: where $\gamma$ can be any graph, $N$ is the number of edges of the graph 
774: $\gamma$ and $f_\gamma:\;SU(2)^N \to \Cl$ is a complex valued function 
775: on $N$ copies of $SU(2)$. \\
776: The holonomy operator acts by multiplication
777: \be \label{0.11}
778: [\widehat{A(e)}\; \psi](A):=A(e)\; \psi(A)
779: \ee
780: while the fluxes act by derivation
781: \be \label{0.12}
782: [\widehat{E_f(S)}\; \psi](A):=i\hbar \{E_f(S),\psi(A)\}
783: \ee
784: The scalar product between functions $\psi,\psi'$ defined via (\ref{0.10})
785: over graphs $\gamma,\gamma'$ respectively is defined as follows:
786: Take the union graph $\gamma^\dprime:=\gamma\cup\gamma'$ and write 
787: the edges $e,e'$ of $\gamma,\gamma'$ respectively as compositions 
788: of edges $e^\dprime$ of $\gamma^\dprime$, say $e=e^\dprime_1\circ ..\circ
789: e^\dprime_n$. Now use the fact that the holonomy factorizes, e.g. 
790: $A(e)=A(e^\dprime_1)..A(e^\dprime_n)$ to write $\psi,\psi'$ as functions 
791: over $\gamma^\dprime$. Then
792: \be \label{0.13}
793: <\psi,\psi'>:=\int_{SU(2)^N}\;d\mu_H(h_1)\;..\;d\mu_H(h_N)\;
794: \overline{\psi_{\gamma^\dprime}(h_1,..,h_N)} \;
795: \psi'_{\gamma^\dprime}(h_1,..,h_N) 
796: \ee
797: where $N$ is the number of edges of $\gamma^\dprime$ and $\mu_H$ is the 
798: Haar measure on $SU(2)$. Specifically, if we write 
799: $h=\cos(\chi) 1_2+\tau_j n^j(\theta,\varphi) \sin(\chi)$ with 
800: $\chi,\theta\in [0,\pi)$ and $\varphi\in [0,2\pi)$ where \\
801: $\vec{n}=(\sin(\theta) 
802: \cos(\varphi),\sin(\theta)\sin(\varphi),\cos(\theta))$ is the usual unit 
803: vector on the sphere then $d\mu_H(h)=c\sin^2(\chi)\sin(\theta) 
804: d\chi\;d\theta\;d\varphi$ and $c$ is a normalization constant fixed 
805: by requiring that $\mu_H(SU(2))=1$.
806: 
807: It turns out that the Hilbert space ${\cal H}_{Kin}$ has a convenient
808: orthonormal basis consisting of the spin network functions (SNF) 
809: $T_{\gamma,\vec{j},\vec{I}}$. For the purpose of this paper 
810: it is enough to say that they are specific functions of the type 
811: (\ref{0.10}) which are polynomials in the holonomies $A(e),\;e\in 
812: E(\gamma)$. They are labelled by a graph $\gamma$, a collection 
813: $\vec{j}:=\{j_e\}_{e\in E(\gamma)}$ of spin quantum numbers 
814: $j_e,\;2j_e=1,2,3,..$, one for each edge, and a collection of other 
815: quantum numbers $\vec{I}=\{I_v\}_{v\in V(\gamma)}$ called intertwiners, 
816: one for each vertex. Since the label $\gamma$ is continuous, the basis is 
817: uncountable and the Hilbert space is therefore not separable. We will 
818: use the short hand $s=(\gamma(s),\vec{j}(s),\vec{I}(s))$ for a spin 
819: network label.
820: 
821: Let us now turn to the quantum constraints. The Gauss constraint imposes 
822: certain restrictions on the quantum numbers $\vec{I}$ of the SNF and is 
823: easily taken care of. The spatial diffeomorphism constraint enforces that
824: we should identify all spin network labels $s,s'$ if there exists a 
825: spatial diffeomorphism $\varphi$ of $\sigma$ such that
826: $\varphi(\gamma(s))=\gamma(s')$ and so is also easily taken care of.
827: The real challenge for LQG (and any other quantum gravity theory) is the 
828: Hamiltonian constraint. Looking at (\ref{0.4}) the mere definition of
829: the operator $\hat{C}$ looks like a hopeless task due to the non -- 
830: linearity of the constraint. It is here where the key technique of 
831: \cite{10} comes into play:\\ 
832: It is easy to see that the nonlinear expression involving the electric 
833: field $E$ can be written in the compact Poisson bracket form
834: \be \label{0.14}
835: e_a^j=\frac{\epsilon_{abc} \epsilon^{jkl} E^b_k E^c_l}{\sqrt{|\det(E)|}}
836: \propto \{A_a^j,V\} \mbox{ where } V=\int_\sigma d^3x \sqrt{|\det(E)|}
837: \ee
838: is the volume\footnote{In the case of $k=0$ this diverges classically.
839: However, we only need its Poisson bracket which is well defined.} of 
840: $\sigma$. The advantage of this way of writing 
841: (\ref{0.14}) is that there exists a volume operarator corresponding to $V$
842: which can be diagonalized in terms of spin network states. Thus, 
843: (\ref{0.14}) can be quantized by substituting Poisson brackets by 
844: commutators divided by $i\hbar$. Next, consider the smeared quantity
845: $C(N)=\int d^3x N(x) C(x)$ where $N$ is a test function. We may write this 
846: as $C(N)=C_F(N)+C_K(N)$ where $C_F,C_K$ respectively correspond to the 
847: terms in (\ref{0.4}) involving the term $F_{ab}^j$ or $K_a^m K_b^n$ 
848: respectively. Now it is easy to see that 
849: \be \label{0.15}
850: K_a^j(x)\propto \{A_a^j(x),\{C_F(1),V\}\}
851: \ee
852: where $C_F(1):=C_F(N)_{N=1}$. It follows that we know how to quantize 
853: $C(N)$ once we know how to quantize $C_F(N)$\footnote{The condition 
854: $C(x)=0$ for all $x$ is equivalent to requiring $C(N)=0$ for all $N$ 
855: because the support of $N$ can be localized arbitrarily.}. In order to do 
856: this it remains to write $C_F(N)$ in terms of holonomies. This is achieved 
857: by recalling that for a curve $e^a_{x,u}(t):=x^a+t u^a$ and loop 
858: $\alpha_{x,u,v}(\epsilon):=
859: e_{x,u}(\epsilon) \circ
860: e_{x+\epsilon u,v}(\epsilon) \circ
861: e_{x+\epsilon(u+v),u}(\epsilon)^{-1} \circ
862: e_{x+\epsilon v,v}(\epsilon)^{-1}$
863: respectively we have the expansions 
864: $A(e_{x,u}(\epsilon))=1_2+\epsilon u^a A^j_a(x) \tau_j/2+O(\epsilon^2)$ 
865: and $A(\alpha_{x,u,v}(\epsilon))=1_2+\epsilon^2 u^a v^b F_{ab}^j(x) 
866: \tau_j/2+O(\epsilon^3)$. 
867: Therefore, if we write the integral $C_F(N)$ as the limit of a Riemann sum 
868: over boxes of coordinate volume $\epsilon^3$ then in the limit it is 
869: appropriate to replace $F_{ab}^j$ by a holonomy along a loop on the 
870: boundary of the box while the $A_a^j$ in the Poisson bracket $\{A_a^j,V\}$
871: may be replaced by a holonomy alon an edge on the boundary of that box.
872: 
873: The details of taking that limit can be inferred from \cite{10,11}.
874: Roughly speaking, one defines (a spatially diffeomorphism invariant 
875: version of) the constraint on the space of solutions to the diffeomorphism
876: constraint on which the limit becomes trivial because what matters to the 
877: constraint are only the terms of the Riemann sum next neighbouring the 
878: vertices and those terms are diffeomorphic to each other when we refine 
879: the Riemann sum. 
880: 
881: \subsection{Elements of LQC and Comparison with LQG}
882: \label{s0.2}
883: 
884: LQC, by definition, is the canonical quantization of the cosmological 
885: truncation of classical GR by the methods of LQG\footnote{Originally 
886: one tried to define LQG as a symmetric sector of distributions on a dense 
887: sunspace of the LQG Hilbert space together with the dual action of the 
888: full LQG operators. However, this dual action does not preserve the 
889: symmetric sector. See also the end of this section. We adopt here the 
890: correct and modern point of view spelled out in \cite{17ba}.}. 
891: Hence one first 
892: specializes the functions $A_a^j(x), E^a_j(x)$ to the spatially 
893: homogeneous case, plugs those into the formulae for the holonomies and
894: fluxes and then computes the corresponding algebra $\mathfrak{A}_{cosmo}$. 
895: That algebra is of course tremendously smaller than the full algebra 
896: $\mathfrak{A}$ but one quatizes it in analogy to LQG, that is, one chooses 
897: a Hilbert space representation which is as similar as 
898: possible to the one of LQG. Finally, one plugs the simplified expression 
899: for $(A,E)$ into the constraints and quantizes them as in LQG, in 
900: particular one uses the technique (\ref{0.14}) which enables to 
901: avoid denominators (which are singular at the classical singularity)
902: at the price of employing Poisson brackets with the volume functional.\\
903: \\
904: Let us see this in more detail\footnote{What follows is a drastically 
905: simplified version of \cite{17ba}.}:\\
906: We will consider the simplest case: The isotropic, flat FRW cosmologies
907: for which the line element reads
908: \be \label{0.16a}
909: ds^2=-dt^2+a(t)^2 d\vec{x}^2
910: \ee
911: where $a$ is the scale factor. Here we take $t,a$ to have dimensions
912: of length.
913: One fixes both the spatial diffeomorphism constraint and the Gauss 
914: constraint by setting
915: \be \label{0.16}
916: A_a^j:=q \delta_a^j/3,\;\;E^a_j:=p\delta^a_j
917: \ee
918: where $p,q$ are spatial constants. It follows that $a=\sqrt{|p|}$ and 
919: that $\{p,q\}=8\pi \beta G$ are canonically conjugate. It is easy to see 
920: that with 
921: (\ref{0.16}) both the Gauss and spatial diffeomorphism constraint vanish
922: identically as they should and that the gravitational 
923: contribution to the Hamilonian constraint becomes 
924: $C=- \alpha q^2 \sqrt{|p|}$ with a positive constant $\alpha$.
925: For a (homogenous) 
926: scalar field we find
927: \be \label{0.17}
928: C^{scalar}\propto \frac{\pi^2}{2\sqrt{|p|}^3}+\sqrt{|p|}^3 V(\phi)
929: \ee
930: As one can show (see the end of section \ref{s4}) the classical 
931: singularity is encoded in the kinetic term only.
932: 
933: The holonomy and flux become respectively 
934: \be \label{0.18}
935: A(e)=\cos(q r_e)+\tau_j n^j_e \sin(r_e q),\;\;       
936: E_f(S)=[\int_S ds_a f^a] p
937: \ee
938: where $e^a(1)-e^a(0)=:r_e n^j_e,\;\;(n_e^j )^2=1$.
939: By varying $e$ and taking traces we may generate the functions 
940: $T_r:=\exp(i r q)$ where $r$ can be any real number. The reduced 
941: holonomy flux algebra therefore is generated by the $T_r$ and $p$.
942: 
943: The reduced Hilbert space is now obtained by asking that the $T_r$ form an 
944: orthonormal basis, i.e. $<T_r,T_{r'}>=\delta_{r,r'}$ so that they play the 
945: analogon of the SNF. Here the continuous graph label and the discrete spin 
946: label of the SNF have joined into one continuous label $r$.
947: The operators $\widehat{T_r},\;\hat{p}$ act by multiplication and 
948: derivation respectively, explicitly $\widehat{T_r}\;T_{r'}=T_{r+r'}$ and 
949: $\hat{p}\;T_r:=i\hbar\{p,T_r\}=-r\ell_p^2\beta  T_r$ where $\ell_p^2=
950: 8\pi G \hbar$ is the Planck length squared.
951: 
952: Finally, in order to quantize objects like the inverse scale 
953: factor $1/a=1/\sqrt{|p|}$ we notice the 
954: classical identity 
955: \be \label{0.19} 
956: \frac{1}{\sqrt{|p|}}\propto \frac{T_{r_0} \{T_{-r_0},\sqrt{|p|}\} -T_{-r_0} 
957: \{T_{r_0},\sqrt{|p|}\}}{i 8 \pi \beta G r_0} 
958: \ee 
959: which holds for every 
960: $r_0\not=0$. Notice that $\sqrt{|p|}\propto V^{1/3}$ where $V$ is the 
961: volume\footnote{For $k=0$ this volume is actually infinite. We mean here 
962: the volume divided by the comoving volume.} so that (\ref{0.19})
963: is indeed analogous to (\ref{0.14}). Notice that the spectrum of the 
964: volume 
965: operator is essentially given by $|r|^{3/2}$ with $T_r$ as normalizable 
966: eigenfunctions (it takes a continuous range but consists only of 
967: eigenvalues, there is no continuous part). Hence the only zero volume 
968: state is $T_0=1$. It follows immediately that if the inverse scale factor 
969: can be densly defined on the $T_s$ then at zero volume it is bounded.  
970: 
971: Expression (\ref{0.19}) is readily quantized by substituting Poisson 
972: brackets by commutators divided by $i\hbar$. This gives
973: \be \label{0.20}
974: \widehat{\frac{1}{\sqrt{|p|}}} T_r 
975: \propto 
976: \frac{T_{r_0} \widehat{\sqrt{|p|}} T_{-r_0} 
977: -T_{-r_0} \widehat{\sqrt{|p|}} T_{r_0}}{\beta \ell_P^2 r_0}\;\;T_r
978: =\frac{\sqrt{|r+r_0|}-\sqrt{|r-r_0|}}{r_0 \sqrt{\beta} \ell_P} \;T_r
979: \ee
980: which is bounded (in fact vanishes) at $r=0$. It is even bounded for all 
981: $r$ and coincides with the operator $1/\sqrt{|\hat{p}|}$ for $r\gg r_0$.
982: The parameter $r_0$ is a quantization ambiguity, called a factor ordering 
983: ambiguity which is also present in LQG \cite{17c}
984: although it is there labelled by a 
985: discrete rather than continuous parameter.  Notice that by the constraint 
986: $C=C_{geo}+C_{scalar}=0$ we obtain $q^2\propto p^2/(2a^4)+V a^2$ which 
987: means that $q\propto a^{-2}$ at $a\to 0$ (see the discussion at the end of 
988: section \ref{s4}) so that $C_{geo}\propto a^{-3}$ classically. However,
989: quantum mechanically $C_{geo}$ is bounded at $a\to 0$ because $q$ is 
990: quantized by the bounded expression $(T_{r_0}-T_{-r_0})/(2i r_0)$. 
991: 
992: So far we have only discussed the isotropic model. In the non -- 
993: isotropic but diagonal model \cite{13a} similar remarks apply because 
994: the three scale factors mutually commute. Hence the inverse scale 
995: factors can be treated separately just as in the isotropic case.\\ 
996: \\
997: Let us summarize:\\
998: LQC is the usual cosmological minisuperspace phase space quantized by 
999: LQG methods. It borrows in an essential way the Hilbert space \cite{5a}
1000: and the key technique \cite{10}. LQC therefore seems to to confirm 
1001: the aspects \cite{5a,10} for LQG. 
1002: 
1003: However, despite these similarities there are 
1004: important differences:\\
1005: To begin with, LQC does not have a graph label. That is of 
1006: course 
1007: expected because in cosmology all points are equivalent due to 
1008: homogeneity, hence there is only one point and thus no room for a 
1009: graph.
1010: Consequently, there is no such thing as the valence of a vertex.
1011: These facts lead to an enormous simplification of the volume spectrum 
1012: which is only modestly degenerate (multiplicity two for all
1013: $r\not=0$, multiplicity one for $r=0$) while in LQG the volume spectrum is 
1014: infinitely degenerate, in particular for zero eigenvalues (just apply a 
1015: diffeomorphism to a given volume eigenstate which changes the graph). 
1016: Moreover, the 
1017: volume spectrum in LQG has discrete range while in LQC it has continuous 
1018: range. Next, in LQG the Hamiltonian constraint modifies the graph of a 
1019: spin network state while in LQC that is not possible since there is no 
1020: graph. That leads to an enormous simplification concerning the 
1021: computation of the kernel of the Hamiltonian constraint which boils down 
1022: to the solution of a recursion relation within a finite 
1023: dimensional parameter space while in LQG the corresponding parameter 
1024: space \cite{10b} is infinite dimensional. 
1025: Finally, while it is possible to interpret the LQC states as 
1026: distributional LQG states \cite{17d} (simply multiply an ordinary LQG 
1027: state by a 
1028: $\delta$ distribution supported on homogeneous connections) this 
1029: distributional space of states is not preserved by the (dual) 
1030: action of the operators of 
1031: LQG, hence LQC is not an invariant (distributional) sector of LQG.
1032: It really is a model whose connection to LQG involves the classical 
1033: truncation, thus inhomogeneous fluctuations are switched off by hand
1034: rather than being quantum mechanically suppressed.\\
1035: \\
1036: We conclude: By definition of a toy model with a finite dimensional phase 
1037: space,
1038: LQC cannot model the QFT aspects of LQG. LQC models very successfully two 
1039: aspects of LQG, namely those of \cite{5a,10}. However, it is far from 
1040: granted that those predictions of LQC which are sensitive to the QFT  
1041: aspects will hold in LQG as well. As we just saw, the spectrum of the 
1042: inverse scale 
1043: factor {\it is} sensitive to those aspects since the volume operator is. 
1044: It is precisely one of the 
1045: purposes of the present paper to investigate whether these differences 
1046: lead to a qualitatively different picture in LQG when applied to 
1047: cosmology.
1048: 
1049: 
1050: \section{Unboundedness of the Analogon of the Inverse Scale Factor in 
1051: Full LQG}
1052: \label{s2}
1053: 
1054: We can explain the reason for the unboundedness of the operator 
1055: concerning the inverse scale factor in very intuitive terms 
1056: without going through the (admittedly lengthy) analysis. We will 
1057: compare this to the isotropic LQC model, similar arguments apply to the 
1058: the diagonal models:\\ 
1059: The spectrum of the inverse scale factor can, roughly speaking, be 
1060: obtained by taking the difference between neighbouring eigenvalues of the 
1061: volume operator (discrete derivative). This is the case because, according 
1062: to (\ref{0.14}) the 
1063: inverse scale factor can be obtained from the Ko -- Dreibein as
1064: $e=\{A,V\}$ where $V$ is the volume and $A$ is the connection. 
1065: This is the 
1066: key technique of \cite{10} as mentioned above. Upon quantization the 
1067: connection gets replaced by a holonomy, that is, $\hat{e}\propto
1068: h[h^{-1},\hat{V}]=V-h\hat{V} h^{-1}$ where $h$ is a the holonomy of $A$ 
1069: along an edge. The holonomy changes the spin labels of a spin network 
1070: state by one unit, hence we get a discrete derivative in the spin 
1071: parameter space. Now the spectrum of the volume operator 
1072: depends parametrically, among 
1073: other things, precisely on the spins of the edges of a spin network 
1074: attached to a vertex and additional intertwiner quantum numbers and this 
1075: is the discrete derivative of the volume spectrum that we mentioned.
1076: 
1077: If we view the spectrum 
1078: of the volume operator as a function of 
1079: these parameters then in isotropic LQC this parameter space is 
1080: one dimensional, it is a 
1081: function $j\mapsto V(j)$ (in LQC the parameter $j:=r$ of the previous 
1082: section takes continuous and not discrete values but that does not change 
1083: the qualitative picture). In LQG on the 
1084: other hand it is arbitrarily high dimensional, the dimensionality being 
1085: governed by the number $n$ of vertices and the valences $N_v$ of the 
1086: individual vertices, it is a function of the form
1087: $(j^{(1)}_1,..,j^{(n)}_{N_n})\mapsto \sum_{v=1}^n 
1088: V_v(j^{(v)}_1,..,j^{(v)}_{N_v})$ where 
1089: the $n,N_v$ can be arbitrarily large. In 
1090: both cases the functions $V$ are 
1091: nowhere singular except when at least one of the $j$ tends to infinity.
1092: 
1093: The point is now that in isotropic LQC the zero volume condition
1094: $V(j)=0$ has a finite number of solutions $j=j_0<\infty$ (in fact one).
1095: The spectrum of 
1096: the inverse scale factor at zero volume is then, roughly speaking, given 
1097: by $V(j_0+\delta j)-V(j_0)$ (using suitable discrete units for the 
1098: parameters $j$, in LQC $\delta j=r_0$) which is 
1099: finite since $j_0<\infty$ whence we get boundedness 
1100: at zero volume. In LQG on the other hand the situation is more 
1101: interesting: The zero volume condition now gives an infinite number
1102: (parameter space of codimension one) of solutions. Pictorially 
1103: speaking, the spectrum 
1104: of the volume operator plotted against the parameter space can be 
1105: envisioned as a complicated landscape of mountains which is disrupted by 
1106: many valleys of zero volume which intersect each other in a complicated 
1107: way. The boundaries of those valleys have arbitrarily steep inclination
1108: and the mountains are arbitrarily high. This is 
1109: because the solutions of the zero volume condition are now of the 
1110: form $j^{(v)}_1=j^{(v)}_0(j^{(v)}_2,..,j^{(v)}_{N_v})$ and we may let 
1111: $j^{(v)}_2,..,j^{(v)}_{N_v}\to\infty$ in many 
1112: ways (along many valleys). It is clear that now for each $v$ 
1113: $V_v((j^{(v)}_1=j^{(v)}_0,j^{(v)}_2,..,j^{(v)}_{N_v}))-
1114: V_v((j^{(v)}_1=j_0,j^{(v)}_2,..,j^{(v)}_{N_v})+\delta j^{(v)})$
1115: with $\delta j^{(v)}_k=\delta_{kl},\;l\in\{1,..,N_v\}$ can become arbitrarily 
1116: large. This is what we are going to prove explicitly in our companion 
1117: paper \cite{16} based on the recent simplification of the analytical
1118: expression for the matrix elements of the volume operator \cite{17}
1119: \footnote{For the very special case of matrix elements we are going to use a similar expression was also derived in \cite{de Pietri} using a different method.}.
1120: 
1121: One may object that the diverging configurations 
1122: $(j^{(v)}_1,j^{(v)}_2,..,j^{(v)}_{N_v})$ for each individual $v$ do not 
1123: look isotropic in general and that one should require isotropy 
1124: as in LQC which perhaps should enforce 
1125: $j^{(v)}_1=..=j^{(v)}_{N_v}$ which could lead to a bound since the 
1126: parameter space is now one dimensional\footnote{Indeed, if we choose 
1127: $j_1=j_2=j_3=j$ in formula 
1128: (\ref{e'(B)^2 Endresultat eichinvarianter 3-Vertex}) then the result is 
1129: bounded in $j$.}. 
1130: However, here we now invoke a second key difference 
1131: between LQC and LQG: LQC is a quantum mechanical toy model of LQG. It can 
1132: be thought of as LQG where all the vertices of a spin network state are to 
1133: be thought of as equivalent due to homogeneity. In other words, LQC is 
1134: like LQG except that we only consider spin network states with a single 
1135: vertex (and a few edges, usually no more than three). LQG on the other 
1136: hand is a field theory and spin network states 
1137: can have an arbitrary number of vertices. Consider now a zero volume spin 
1138: network 
1139: state $T_{\lambda,v}$ where $\lambda$ is the eigenvalue of the inverse 
1140: scale factor and $v$ is a vertex at which the graph $\gamma_v$ with 
1141: a single non co -- planar\footnote{The volume operator vanishes on 
1142: (non) gauge invariant 
1143: vertices which are not at least (tri) four -- valent and at which not at 
1144: least 
1145: three tangents of the adjacent edges are linearly independent.} 
1146: vertex $v$ of the spin network 
1147: state is located. Let $S$ be a finite\footnote{This will be sufficient
1148: if the spatial manifold is compact as for $k=1$ FRW models or if we are 
1149: interested only a compact, observationally accessible (inside the Hubble 
1150: radius) region. In the 
1151: $k=0,-1$ case we can use in the infinite tensor product \cite{19}.} 
1152: set of $n$ vertices obtained from a random 
1153: sprinkling of points into the spatial manifold \cite{18} and consider the 
1154: state $\psi:=\otimes_{v\in S} T_{\lambda,v}$. 
1155: We may arrange that the $\gamma_v$ are mutually 
1156: disjoint 
1157: and that their orientation is random as well. This state is 
1158: homogeneous and isotropic on large scales but not on microscopic ones 
1159: (the violation scale of homogeneity and isotropy depends on the 
1160: number of points in $S$ and decreases with $n$). The state is normalized 
1161: and still an eigenvalue $\lambda$ eigenstate of the inverse scale factor.
1162: This proves that we may construct kinematical states with large scale 
1163: homogeneity and isotropy in LQG with arbitrarily large norm of the inverse 
1164: scale factor\footnote{This is a purely kinematical result as in LQC.
1165: In fact it must be because the inverse scale 
1166: factor itself is not a Dirac observable, see next section.}.\\
1167: \\
1168: Let us now go into more details:\\
1169: In order to match the LQC calculations as closely as possible, what we 
1170: really have studied is the contribution of a scalar 
1171: (or inflaton) field to the Hamiltonian constraint (smeared energy density)
1172: operator whose classical expression is 
1173: \be \label{3.1}
1174: 2C_{scalar}(N)=\int_\sigma\; d^3x\; N\;[\frac{\pi^2}{\sqrt{\det(q)}}+
1175: \sqrt{\det(q)} (q^{ab} \phi_{,a}\phi_{,b}+2V(\phi))]
1176: \ee
1177: Here $\phi$ is the scalar field, $\pi$ its conjugate momentum,
1178: $q_{ab}$ is the pull back to the spatial manifold $\sigma$ of the four 
1179: metric $g_{\mu\nu}$ on the four manifold $M\cong \Rl\times \sigma$ and 
1180: $N$ is a test function (see also section \ref{s0}). In an exactly 
1181: homogeneous situation, say the FRW
1182: metric, the derivative term drops out and the potential term becomes 
1183: proportional to $a^3$ while the kinetic term becomes proportional to 
1184: $a^{-3}$ where $a$ is the scale factor of the FRW metric. The 
1185: classical initial singularity $\lim_{t\to 0} a(t)=0$ is therefore 
1186: prominent in the kinetic term. Of course in full LQG the kinetic term must 
1187: be and can be considered as well, see \cite{10b}, but it will be 
1188: sufficient to consider the kinetic term to demonstrate our point.
1189: 
1190: Consider therefore the kinetic term
1191: \be \label{3.2}
1192: 2C^{kin}_{scalar}(N)=\int_\sigma\; d^3x\; N\;\frac{\pi^2}{\sqrt{\det(q)}}
1193: \ee
1194: Its quantization is derived in \cite{10b} on the kinematical Hilbert
1195: space ${\cal H}_{kin}={\cal H}_{geo}\otimes {\cal H}_{scalar}$.
1196: An orthonormal basis in this Hilbert space \cite{10d} consists 
1197: of the 
1198: charge -- spin network states $T_{\gamma,\vec{j},\vec{I}}\otimes 
1199: T_{V(\gamma),\vec{n}}$. Here $\gamma$ is a graph with vertex set 
1200: $V(\gamma)$, $\vec{j}$ denotes a labelling of each edge $e$ of $\gamma$ by 
1201: half integral spin quantum numbers $j_e$, and $\vec{I},\;\vec{n}$ 
1202: respectively denote a labelling 
1203: of each vertex $v$ of $\gamma$ with a gauge invariant intertwiner $I_v$ 
1204: and integer $n_v$ respectively. The action of the corresponding 
1205: operator, densely defined in the finite linear span of this basis, 
1206: can now schematically be written as 
1207: \be \label{3.3}
1208: 2\widehat{C^{kin}_{scalar}}(N)\;T_{\gamma,\vec{j},\vec{I},\vec{n}}=
1209: \sum_{v\in V(\gamma)}\; N(v)\;[\widehat{\frac{1}{\sqrt{\det(q)}}(v)}\;
1210: T_{\gamma,\vec{j},\vec{I}}]\;\otimes\;[\widehat{\pi}(v)^2\;
1211: T_{V(\gamma),\vec{n}}]
1212: \ee
1213: The explicit meaning and action of the genuine operators ({\it not} 
1214: operator valued distributions) can be found in \cite{10b,16a} and is 
1215: also reviewed in \cite{16} for the convenience of the reader. For the 
1216: purpose of this paper it is enough to report the result of our 
1217: calculation. 
1218: 
1219: The analogon of the LQC inverse scale factor (cubed) in LQG is of course 
1220: the non -- negative operator $\widehat{\frac{1}{\sqrt{\det(q)}}(v)}$. 
1221: To show that it is unbounded from above on zero volume eigenstates it 
1222: is sufficient to construct one single system of states whose members 
1223: have arbitrarily large norm. The simplest 
1224: zero volume eigenstates are spin network states 
1225: $T_{\gamma,\vec{j},\vec{I}}$ which have at most three valent vertices.
1226: They are very simple because the intertwiners $I_v$ are then uniquely 
1227: determined. We are thus interested in the numbers
1228: \be \label{3.4}
1229: ||\widehat{\frac{1}{\sqrt{\det(q)}}(v)}\;T_{\gamma,\vec{j}}||^2
1230: =<T_{\gamma,\vec{j}},[\widehat{\frac{1}{\sqrt{\det(q)}}(v)}]^2
1231: \;T_{\gamma,\vec{j}}>
1232: \ee
1233: where $v$ is a three valent vertex of $\gamma$ whose adjacent, non co -- 
1234: planar edges carry the spin quantum numbers $j_1 \le j_2 \le j_3$ subject 
1235: to the constraint that $j_3\in \{j_1+j_2,j_1+ j_2-1,..,|j_1-j_2|\}$. To 
1236: make things definite, consider a graph $\gamma$ with three edges and two 
1237: three valent, non co -- planar vertices, see figure \ref{fig1}.\\
1238: %
1239: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1240: % Figure 1
1241: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1242: 
1243: \begin{figure}[hbt]
1244:     \label{fig1}
1245:     \center
1246:     \cmt{7}{
1247:     \psfrag{v1}{${v_1}$}
1248:     \psfrag{v2}{${v_2}$}
1249:     \psfrag{e1}{$e_1$}
1250:     \psfrag{e2}{$e_2$}
1251:     \psfrag{e3}{$e_3$}
1252:     \includegraphics[height=3cm]{pictures/3_valent_graph_with_2_vertices.eps} 
1253:     \caption{Simplest 3 valent gauge invariant graph $\gamma$: $V(\gamma)=\{v_1,v_2\},E(\gamma)=\{e_1,e_2,e_3\}$} }
1254: \end{figure}
1255: 
1256: 
1257: %
1258: 
1259: Remarkably the number (\ref{3.4}) can be computed analytically. This is 
1260: possible due to a recent improvement in the technology for the computation 
1261: of 
1262: the matrix elements of the volume operator for full LQG \cite{17}.
1263: The analytical formula for (\ref{3.4}) is given by the surprisingly 
1264: compact expression\\
1265: %
1266: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1267: % Main formula
1268: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1269: %
1270: \fcmt{18}{\[ \begin{array}{llll}
1271:      \\
1272:     ||\widehat{\frac{1}{\sqrt{\det(q)}}(v)}\;T_{\gamma,\vec{j}}||^2
1273:    &=&\multicolumn{2}{l}{\displaystyle\frac{C^2}{(\ell_P)^6}\cdot\Bigg[
1274:     \displaystyle\frac{32\,}{9\,
1275:     {\left( 1 + 2\,{j_1} \right) }^2\,{\left( 1 + 2\,{j_2} \right) }^2\,{\left( 1 + 2\,{j_3} \right) }^2}}
1276:     
1277:     \\\\
1278:     &&\times&\big[ 108\,A_1 \,A_2 \,
1279:     A_3
1280:       
1281:     \\
1282:     &&&
1283:    - 3\,{\left( 2\,{\left( -1 \right) }^{2\,{j_1}} + 
1284:            {\left( -1 \right) }^{2\,{j_3}} \right) }^2\,A_2 \,
1285:        {\left( A_1 -A_2 + A_3 \right) }^2 
1286:       
1287:     \\
1288:     &&& 
1289:        
1290:        -3\,{\left( 2\,{\left( -1 \right) }^{2\,{j_2}} + {\left( -1 \right) }^{2\,{j_3}} \right) }^2\,
1291:        
1292:        A_1  \,{\left( -A_1 + A_2 + A_3 \right) }^2
1293:        
1294:     \\
1295:     &&&
1296:       
1297:        -3\,{\left( 1 + 2\,{\left( -1 \right) }^{2\,\left( {j_1} + {j_2} \right) } \right) }^2\,
1298:        A_3 \,{\left( A_1 + A_2 - 
1299:            A_3  \right) }^2 \ 
1300:     \\
1301:     &&&  
1302:        -~ \left( 1 + 2\,{\left( -1 \right) }^{2\,\left( {j_1} + {j_2} \right) } \right) \,
1303:        \left( 2\,{\left( -1 \right) }^{2\,{j_1}} + {\left( -1 \right) }^{2\,{j_3}} \right) \,
1304:        \left( 2\,{\left( -1 \right) }^{2\,{j_2}} + {\left( -1 \right) }^{2\,{j_3}} \right) \,
1305:     \\
1306:     &&&~~~~~~~\times
1307:        \left( -A_1 +A_2 +A_3 \right) \,
1308:        \left( A_1 -A_2 + A_3 \right) \,
1309:        \left( A_1 + A_2 - A_3  \right)  
1310:        
1311:      \big]  
1312:      \\\\&&\times&	   
1313:      \,
1314:     {\left( {V_{1A}}^{\frac{1}{4}} - {V_{1B}}^{\frac{1}{4}} \right) }^2\,
1315:     {\left( {V_{2A}}^{\frac{1}{4}} - {V_{2B}}^{\frac{1}{4}} \right) }^2\,
1316:     {\left( {V_{3A}}^{\frac{1}{4}} - {V_{3B}}^{\frac{1}{4}} \right) }^2\Bigg]^2
1317: 
1318: \end{array}\] \be\label{e'(B)^2 Endresultat eichinvarianter 3-Vertex} \ee }
1319: 
1320: 
1321: ~\\\\
1322: where $C$ is a positive  numerical constant depending on the regularization of the operator 
1323: $\widehat{\frac{1}{\sqrt{\det(q)}}(v)}$ and the volume operator.
1324: Furthermore $(j_1,j_2,j_3)$ are the spins of the 3 edges $(e_1,e_2,e_3)$ outgoing at the vertex $v$ and $A_K:=j_K(j_K+1)$. Moreover
1325: 
1326: 
1327: \[\begin{array}{lcl}
1328:     V_{1A}
1329:     &=&\big[(-j_1+j_2+j_3+1)(j_1-j_2+j_3)(j_1+j_2-j_3)(j_1+j_2+j_3+1) \big]^{\frac{1}{2}}
1330:     
1331:     \\
1332:     V_{1B}
1333:     &=&\big[(-j_1+j_2+j_3)(j_1-j_2+j_3+1)(j_1+j_2-j_3+1)(j_1+j_2+j_3+2) \big]^{\frac{1}{2}}
1334: 			    
1335:     \\
1336:     
1337:     V_{2A}
1338:     &=&\big[(-j_1+j_2+j_3)(j_1-j_2+j_3+1)(j_1+j_2-j_3)(j_1+j_2+j_3+1) \big]^{\frac{1}{2}}
1339:     
1340:     
1341:     \\
1342:    
1343:    V_{2B}
1344:        &=&\big[(-j_1+j_2+j_3)(j_1-j_2+j_3)(j_1+j_2-j_3+1)(j_1+j_2+j_3+2) \big]^{\frac{1}{2}}
1345:  
1346:     
1347:     \\
1348:     
1349:     V_{3A}
1350:         &=&\big[(-j_1+j_2+j_3)(j_1-j_2+j_3)(j_1+j_2-j_3+1)(j_1+j_2+j_3+1) \big]^{\frac{1}{2}}
1351: 			    
1352:     \\			    
1353:     V_{3B}
1354:         &=&\big[(-j_1+j_2+j_3+1)(j_1-j_2+j_3+1)(j_1+j_2-j_3)(j_1+j_2+j_3+2) \big]^{\frac{1}{2}}
1355: 		    
1356:     			    
1357: \end{array}\]
1358: 
1359: We now display a few diverging series of configurations $(j_1,j_2,j_3)$ where 
1360: we  always plot \linebreak
1361: $||\widehat{\frac{1}{\sqrt{\det(q)}}(v)}\;T_{\gamma,\vec{j}}||:=\sqrt{||\widehat{\frac{1}{\sqrt{\det(q)}}(v)}\;T_{\gamma,\vec{j}}||^2}$, neglecting the prefactor $\frac{C}{(\ell_P)^3)}$\\
1362: %
1363: \pagebreak
1364: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1365: % Diverging Configurations
1366: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1367: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1368: \subsubsection*{''Oscillating''}
1369: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1370: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1371: \paragraph*{\fbox{$j_1=j_2=\frac{j_3}{2}$}}%
1372: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1373: If we set $j_1=j_2=\frac{j_3}{2}$ where $j_3 \in \mb{N}$, we get:
1374: 
1375: \begin{footnotesize}
1376: \be
1377: ||\widehat{\frac{1}{\sqrt{\det(q)}}(v)}\;T_{\gamma,\vec{j}}||
1378:    \propto
1379: \frac{128\,{\sqrt{2}}\,\left( -1 + {\left( -1 \right) }^{{j_3}} \right) \,{{j_3}}^4\,
1380:     \left( -3\,\left( 2 + {\left( -1 \right) }^{{j_3}} \right)  + 
1381:       \left( 1 + {\left( -1 \right) }^{3\,{j_3}} \right) \,{j_3} \right) }{9\,\left( 1 + {j_3} \right) \,
1382:     {\left( 1 + 2\,{j_3} \right) }^{\frac{7}{4}}}
1383: \ee
1384: 
1385: \begin{figure}[htbp!]        
1386:   \begin{minipage}[t]{7.5cm}
1387:   \psfrag{e2}{\tiny\hspace{-5mm}$\small\|\widehat{\frac{1}{\sqrt{\det(q)}}(v)}\;T_{\gamma,\vec{j}}\|$}
1388:   \psfrag{j3}{$j_3$}
1389:   \includegraphics[height=4.5cm]{pictures/unbound_config1.eps}
1390:   \caption{Plot for $j_1=j_2=\frac{j_3}{2}$ where $j_3 \in \mb{N}$ with $1\le j_3 \le 40$. The graph oscillates between 0 (if $j_3$ even) and an increasing value (if $j_3$ odd)} 
1391:   \end{minipage}
1392:  %\hfill
1393:  \begin{minipage}[b]{8cm}
1394:  Asymtotically this increases as
1395:   \[\begin{array}{lcl}
1396:      ||\widehat{\frac{1}{\sqrt{\det(q)}}(v)}\;T_{\gamma,\vec{j}}||
1397:     & \propto&
1398:     \frac{128\cdot 6 \cdot\sqrt{2} \cdot {j_3}^4}{9(1+j_3)(1+2j_3)^{\frac{11}{4}}}\nonumber\\
1399:     &\stackrel{j_3 \rightarrow \infty}{\propto}&
1400:     35.9 \cdot {j_3}^{\frac{5}{4}}
1401:   \end{array}\]
1402:   \vfill
1403:   \end{minipage}
1404: \end{figure}
1405: 
1406: \end{footnotesize}
1407: 
1408: 
1409: 
1410: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1411: \paragraph*{\fbox{$j_1=j_2=\frac{j_3}{2}+\frac{1}{2}$}}
1412: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1413: \begin{samepage}
1414: 
1415: If we set $j_1=j_2=\frac{j_3}{2}+\frac{1}{2}$ where $j_3 \in \mb{N}$, we get:
1416: 
1417: \begin{footnotesize}
1418: \be\begin{array}{lcl}
1419: ||\widehat{\frac{1}{\sqrt{\det(q)}}(v)}\;T_{\gamma,\vec{j}}||
1420:    &\propto&
1421:     \frac{1}{9\,
1422:     {\left( 2 + {j_3} \right) }^4\,{\left( 1 + 2\,{j_3} \right) }^2}
1423:     
1424:    \\\\
1425:    &&\times~    
1426:     128\,{\sqrt{2}}\,\left( 1 + {\left( -1 \right) }^{{j_3}} \right) \,
1427:     {{j_3}}^2\,{\left( 1 + {j_3} \right) }^3\,\left( -3\,
1428:        \left( -7 + 3\,{\left( -1 \right) }^{{j_3}} + {\left( -1 \right) }^{3\,{j_3}} \right)  + 
1429:       \left( -1 + {\left( -1 \right) }^{3\,{j_3}} \right) \,{j_3} \right) \,
1430:     \\
1431:     &&\times
1432:     {\left( -\left( 2^{\frac{1}{4}}\,{\left( {{j_3}}^2\,\left( 1 + {j_3} \right)  \right) }^{\frac{1}{8}} \right)
1433:             + {\left( {\left( 1 + {j_3} \right) }^2\,\left( 3 + 2\,{j_3} \right)  \right) }^{\frac{1}{8}} \right)
1434:         }^2\,
1435: 	
1436:     \\&&
1437:     \times{\left( {\left( {j_3}\,{\left( 1 + {j_3} \right) }^2 \right) }^{\frac{1}{8}} - 
1438:         {\left( {j_3}\,\left( 3 + 5\,{j_3} + 2\,{{j_3}}^2 \right)  \right) }^{\frac{1}{8}} \right) }^4	
1439: \end{array}\ee
1440: \vspace{-1cm}
1441: \begin{figure}[hbt]
1442:          \begin{minipage}[t]{7.5cm}
1443: 	 \psfrag{e2}{\tiny\hspace{-5mm}$\small\|\widehat{\frac{1}{\sqrt{\det(q)}}(v)}\;T_{\gamma,\vec{j}}\|$}
1444: 	 \psfrag{j3}{$j_3$}
1445: 	 \includegraphics[height=4.5cm]{pictures/unbound_config2.eps}
1446:           \caption{Plot for $j_1=j_2=\frac{j_3+1}{2}$ where $j_3 \in \mb{N}$ with $1\le j_3 \le 40$. The graph oscillates between 0 (if $j_3$ odd) and an increasing value (if $j_3$ even)} 
1447:          \end{minipage}
1448:          \begin{minipage}[b]{8cm}
1449:             Asymtotically this increases as
1450:             \[\begin{array}{lcl}
1451:                 ||\widehat{\frac{1}{\sqrt{\det(q)}}(v)}\;T_{\gamma,\vec{j}}||
1452:                 & \propto&5.9\cdot 10^{-5}~{j_3}^{\frac{5}{4}}
1453:             \end{array}\] 
1454:          \end{minipage}
1455: \end{figure}   
1456: 
1457: \end{footnotesize}   
1458: \end{samepage}
1459: 
1460: 
1461: 
1462: 
1463: \subsubsection*{Increasing}
1464: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1465: \paragraph*{\fbox{$j_1=\frac{3}{2}~ j_2=j_3+\frac{1}{2}$}}%
1466: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1467: If we set $j_1=\frac{3}{2}$, $j_2=j_3+\frac{1}{2}$ where $j_3 \in \mb{N}$, we get:
1468: \begin{footnotesize}
1469: \[\begin{array}{lcl}
1470: ||\widehat{\frac{1}{\sqrt{\det(q)}}(v)}\;T_{\gamma,\vec{j}}||
1471:    &\propto&\displaystyle
1472: 
1473:    	\frac{1}{3\,\left( 1 + {j_3} \right) \,
1474:     {\left( 1 + 2\,{j_3} \right) }^2}
1475:    \\&&\times
1476:    4\,{j_3}\,\left( -9 + 21\,{j_3} + 14\,{{j_3}}^2 \right) \,
1477:     {\left( -\left( 3^{\frac{1}{8}}\,{\left( {j_3}\,\left( 2 + {j_3} \right)  \right) }^{\frac{1}{8}} \right)
1478:             + {\left( -3 + 4\,{j_3} + 4\,{{j_3}}^2 \right) }^{\frac{1}{8}} \right) }^2
1479:     \\&&\times\,
1480:     {\left( -2\,{\left( {j_3}\,\left( 2 + {j_3} \right)  \right) }^{\frac{1}{8}} + 
1481:         {\sqrt{2}}\,3^{\frac{1}{8}}\,{\left( -3 + 4\,{j_3} + 4\,{{j_3}}^2 \right) }^{\frac{1}{8}} \right) }^2\,
1482:     {\left( {\left( {j_3}\,\left( 3 + 2\,{j_3} \right)  \right) }^{\frac{1}{8}} - 
1483:         {\left( -6 + 9\,{j_3} + 6\,{{j_3}}^2 \right) }^{\frac{1}{8}} \right) }^2
1484: 	
1485: 	
1486: 
1487: \end{array}\]
1488: 
1489: \begin{figure}[htbp!]
1490:          \begin{minipage}[t]{7.5cm}
1491: 	    \psfrag{e2}{\tiny\hspace{-5mm}$\small\|\widehat{\frac{1}{\sqrt{\det(q)}}(v)}\;T_{\gamma,\vec{j}}\|$}
1492: 	    \psfrag{j3}{$j_3$}	 
1493: 	    \includegraphics[height=5cm]{pictures/unbound_config3.eps}
1494:               \caption{Plot for $j_1=\frac{3}{2}$, $j_2=j_3+\frac{1}{2}$ 
1495:  where $j_3 \in \mb{N}$ with $1\le j_3 \le 40$. The graph first decreases 
1496: for $1\le j_3 <3 $ and is 0 for $j_3=3$ . It increases for $j_3>3$}
1497: 	  \end{minipage}  
1498:           \begin{minipage}[b]{8cm}
1499:              Asymtotically this increases as
1500:              \[\begin{array}{lcl}
1501:                  ||\widehat{\frac{1}{\sqrt{\det(q)}}(v)}\;T_{\gamma,\vec{j}}||
1502:                  & \propto& 1.06\cdot 10^{-6}~{j_3}^{\frac{3}{2}}
1503:              \end{array}\]
1504:           \end{minipage}
1505: \end{figure}
1506: \end{footnotesize}
1507: 
1508: %
1509: 
1510: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1511: \subsubsection*{\label{Nullkonfigurationen}Identical 0}
1512: If we set $j_1=j_2=j_3$ and more general $j_1,j_2,j_3\in \mb{N}$ (all spins integer numbers) in (\ref{e'(B)^2 Endresultat eichinvarianter 3-Vertex})  then 
1513: 
1514: \[\begin{array}{lcl}
1515: ||\widehat{\frac{1}{\sqrt{\det(q)}}(v)}\;T_{\gamma,\vec{j}}||
1516:    &=&0
1517: 
1518: \end{array}\] 
1519: \vfill
1520: 
1521: 
1522: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1523: \subsection*{General Configurations}
1524: 
1525: Finally we plot the numbers (\ref{3.4}) as $j_1,j_2$ vary over their 
1526: allowed values at given $j_3$ for various values of $j_3$. The landscape of low valleys together with their arbitrarily high and steep boundaries is nicely illustrated:\\
1527: 
1528: Taking the general result (\ref{e'(B)^2 Endresultat eichinvarianter 3-Vertex}) (without the prefactors $(\ell_P)^9|Z|^\frac{3}{2}$)  we use the quantity
1529: 
1530: \[
1531:   Q=\left\{\begin{array}{lcl}
1532:                              30+\ln{[||\widehat{\frac{1}{\sqrt{\det(q)}}(v)}\;T_{\gamma,\vec{j}}||
1533: 			     (j_1,j_2,j_3)]}&&|j_1-j_2|\le j_3 \le j_1+j_2~~and~~j_1+j_2+j_3~is~integer~~\\
1534: 			     &&and~~ ||\widehat{\frac{1}{\sqrt{\det(q)}}(v)}\;T_{\gamma,\vec{j}}||(j_1,j_2,j_3)\ne 0
1535: 			     \\\\
1536: 			     0 &~~~&else
1537:                          \end{array}\right.
1538: \]
1539: where the 30 is arbitrarily added in order to shift the plot upwards.
1540: We make a three dimensional plot (in the range $\frac{1}{2}\le j_1 \le j_{max}$, $\frac{1}{2}\le j_2 \le j_{max}$ for each fixed value $5\le j_3 \le \frac{15}{2}$ :
1541: 
1542: I turns out that the non-zero configurations are grouped symetrically along lines parallel to the $j_1=j_2$ axis. The reason for this is of course the integer requirement $j_1+j_2+j_3\stackrel{!}{=}integer$.
1543: Therefore we will get contributions on the $j_1=j_2$-axis only if $j_3$ is integer.
1544: Because (\ref{e'(B)^2 Endresultat eichinvarianter 3-Vertex}) is symmetric with respect to the interchange of $j_1 \leftrightarrow j_2$ we may restrict  ourselves to the range $j_1\ge j_2$. 
1545: 
1546: 
1547: 
1548: Additionally we contribute for each of those 3D-plot a 2 dimensional plot along the lines
1549: 
1550: \[
1551:   j_2=j_1 - l~~~~~~\mbox{mit}~~0\le l\le \min[j_3,j_{max}-\frac{1}{2}]~,~~~ l+\frac{1}{2}\le j_1 \le j_{max}=25
1552: \]
1553: 
1554: The restriction for the parameter $l$ is a result from the requirements $|j_1-j_2|\le j_3 \le j_1+j_2$ from which for $j_1>j_2$ we may remove the modulus. 
1555:  In order to give a better impression we have joined only non-vanishing 
1556:  values of $Q$ along the lines described above\footnote{For integer $j_3$ 
1557:  also every second configuration on the lines gives a 0 (the identically
1558: $0$ case). Hence, if we would join all the points, the plots would 
1559:  oscillate between 0 and the now plotted curves and it would be hardly 
1560: possible to 
1561: see anything useful from them.}.
1562: 
1563: 
1564:  \clearpage
1565:  \begin{figure}[htbp!]
1566:          \cmt{8}{
1567: 	 \psfrag{j_1}{$j_1$}
1568: 	 \psfrag{j_2}{$j_2$}
1569: 	 \psfrag{ln[eB^2]}{$Q$}
1570: 	 \psfrag{j_3=5}{}
1571: 	 \includegraphics[height=6cm]{pictures/j3is5_3D.eps}
1572:           \caption{Plot for $j_3=5$} }
1573: 	 \cmt{8}{
1574: 	 \psfrag{j_1}{$j_1$}
1575: 	 \psfrag{j_2}{$j_2$}
1576: 	 \psfrag{ln[eB^2]}{$Q$}
1577: 	 \psfrag{j_3=5}{}
1578: 	 
1579: 	 \includegraphics[height=6cm]{pictures/j3is5_2D.eps}
1580:           \caption{Plot for $j_3=5$} }
1581:  
1582: \end{figure}
1583: 
1584: 
1585: \begin{figure}[htbp!]
1586:          \cmt{8}{
1587: 	 \psfrag{j_1}{$j_1$}
1588: 	 \psfrag{j_2}{$j_2$}
1589: 	 \psfrag{ln[eB^2]}{$Q$}
1590: 	 \psfrag{j_3=11/2}{}
1591: 	 
1592: 	 \includegraphics[height=6cm]{pictures/j3is11_2_3D.eps}
1593:           \caption{Plot for $j_3=\frac{11}{2}$} }
1594: 	 \cmt{8}{
1595: 	 \psfrag{j_1}{$j_1$}
1596: 	 \psfrag{j_2}{$j_2$}
1597: 	 \psfrag{ln[eB^2]}{$Q$}
1598: 	 \psfrag{j_3=11/2}{}
1599: 	 \includegraphics[height=6cm]{pictures/j3is11_2_2D.eps}
1600:           \caption{Plot for $j_3=\frac{11}{2}$, $j_2=j_1-l$. The different curves are (bottom to top):
1601: 	  $l=\frac{1}{2},\frac{3}{2},\ldots,\frac{11}{2}$} }
1602:  
1603: \end{figure}
1604: 
1605: 
1606: \begin{figure}[htbp!]
1607:          \cmt{8}{
1608: 	 \psfrag{j_1}{$j_1$}
1609: 	 \psfrag{j_2}{$j_2$}
1610: 	 \psfrag{ln[eB^2]}{$Q$}
1611: 	 \psfrag{j_3=6}{}
1612: 	 \includegraphics[height=6cm]{pictures/j3is6_3D.eps}
1613:           \caption{Plot for $j_3=6$} }
1614: 	 \cmt{8}{
1615: 	 \psfrag{j_1}{$j_1$}
1616: 	 \psfrag{j_2}{$j_2$}
1617: 	 \psfrag{ln[eB^2]}{$Q$}
1618: 	 \psfrag{j_3=6}{}
1619: 	 \includegraphics[height=6cm]{pictures/j3is6_2D.eps}
1620:           \caption{Plot for $j_3=6$} }
1621:  
1622: \end{figure}
1623: 
1624: \begin{figure}[htbp!]
1625:          \cmt{8}{
1626: 	 \psfrag{j_1}{$j_1$}
1627: 	 \psfrag{j_2}{$j_2$}
1628: 	 \psfrag{ln[eB^2]}{$Q$}
1629: 	 \psfrag{j_3=13/2}{}
1630: 	 \includegraphics[height=6cm]{pictures/j3is13_2_3D.eps}
1631:           \caption{Plot for $j_3=\frac{13}{2}$} }
1632: 	 \cmt{8}{
1633: 	 \psfrag{j_1}{$j_1$}
1634: 	 \psfrag{j_2}{$j_2$}
1635: 	 \psfrag{ln[eB^2]}{$Q$}
1636: 	 \psfrag{j_3=13/2}{}
1637: 	 \includegraphics[height=6cm]{pictures/j3is13_2_2D.eps}
1638:           \caption{Plot for $j_3=\frac{13}{2}$, $j_2=j_1-l$. The different curves are (bottom to top):
1639: 	  $l=\frac{1}{2},\frac{3}{2},\ldots,\frac{13}{2}$} }
1640:  \end{figure}
1641: 
1642: 
1643: \begin{figure}[htbp!]
1644:          \cmt{8}{
1645: 	 \psfrag{j_1}{$j_1$}
1646: 	 \psfrag{j_2}{$j_2$}
1647: 	 \psfrag{ln[eB^2]}{$Q$}
1648: 	 \psfrag{j_3=7}{}
1649: 	 \includegraphics[height=6cm]{pictures/j3is7_3D.eps}
1650:           \caption{Plot for $j_3=7$} }
1651: 	 \cmt{8}{
1652: 	 \psfrag{j_1}{$j_1$}
1653: 	 \psfrag{j_2}{$j_2$}
1654: 	 \psfrag{ln[eB^2]}{$Q$}
1655: 	 \psfrag{j_3=7}{}
1656: 	 \includegraphics[height=6cm]{pictures/j3is7_2D.eps}
1657:           \caption{Plot for $j_3=7$} }	   
1658: \end{figure}	
1659: 
1660: 
1661: \begin{figure}[htbp!]
1662:          \cmt{8}{
1663: 	 \psfrag{j_1}{$j_1$}
1664: 	 \psfrag{j_2}{$j_2$}
1665: 	 \psfrag{ln[eB^2]}{$Q$}
1666: 	 \psfrag{j_3=15/2}{}
1667: 	 \includegraphics[height=6cm]{pictures/j3is15_2_3D.eps}
1668:           \caption{Plot for $j_3=\frac{15}{2}$} }
1669: 	 \cmt{8}{
1670: 	 \psfrag{j_1}{$j_1$}
1671: 	 \psfrag{j_2}{$j_2$}
1672: 	 \psfrag{ln[eB^2]}{$Q$}
1673: 	 \psfrag{j_3=15/2}{}
1674: 	 \includegraphics[height=6cm]{pictures/j3is15_2_2D.eps}
1675:           \caption{Plot for $j_3=\frac{15}{2}$, $j_2=j_1-l$. The different curves are (bottom to top):
1676: 	  $l=\frac{1}{2},\frac{3}{2},\ldots,\frac{15}{2}$} } 
1677: \end{figure}
1678: ~\\
1679: 
1680: \pagebreak 
1681: Due to the large kernel of the volume operator even for four -- and higher 
1682: valent vertices \cite{17} a qualitatively similar behaviour will prevail 
1683: in those cases as well. The unboundedness of the ``inverse scale factor''
1684: operator in full LQG, even on homogeneous and isotropic states is herewith 
1685: established: Let $\gamma_1,..,\gamma_n$ be a random distribution of
1686: mutually disjoint but diffeomorphic 
1687: graphs, say of the type of figure \ref{fig1} which is homogeneous and 
1688: isotropic with respect to the comoving spatial metric of a FRW spacetime 
1689: on a scale large compared to the Planck length within a compact 
1690: region (the smaller that scale the larger $n$). The combined graph 
1691: $\gamma:=\cup_{k=1}^n\; \gamma_k$ looks then homogeneous and 
1692: isotropic. Consider the state of unit norm 
1693: \be \label{3.5}
1694: T_{\gamma,\lambda}:=\otimes_{k=1}^n \; T_{\gamma_k,\vec{j}} 
1695: \ee
1696: where we choose the same triple of spins for every graph and such that the 
1697: numbers (\ref{3.4}) are equal to the same value $\lambda^2$ for every 
1698: vertex of the $\gamma_k$. Then obviously all the norms (\ref{3.4}) with 
1699: $T_{\gamma,\vec{j}}$ chosen as the homogeneous state (\ref{3.5}) equal 
1700: $\lambda^2$ for every of the $2n$ vertices of $\gamma$. But $\lambda$ can 
1701: be arbitrarily large. \\
1702: \\
1703: Let us compare this with the situation for the hydrogen atom: In both 
1704: LQG and LQC
1705: the spectrum of the inverse scale factor is purely discrete, so we should 
1706: compare with the bound states of the hydrogen atom. Restricted to the 
1707: subspace of bound states it is indeed the case that the Hamiltonian of the 
1708: hydrogen atom is bounded from below and above and this is not changed 
1709: when we switch from the finite number of degrees of freedom of QM to the 
1710: infinite number of degrees of freedom of QED. However, the interplay 
1711: between LQC and LQG is different: In LQC we have boundedness, but after 
1712: switching 
1713: on LQG we get unboundedness.
1714: 
1715: 
1716: 
1717: \section{Boundedness of the Inverse Scale Factor Expectation Values at the 
1718: Initial Singularity}
1719: \label{s4}
1720: 
1721: 
1722: The result of the previous section sounds rather negative at first. 
1723: However, it is not as we will
1724: now explain. It just shows that {\bf the boundedness of the inverse scale 
1725: factor cannot be a mechanism for avoidance of the local singularity 
1726: in LQG.} 
1727: 
1728: Namely, what
1729: we have shown is unboundedness on certain homogeneous and isotropic 
1730: spin network states. However, that does not 
1731: mean that the inverse scale factor is unbounded on all homogeneous 
1732: and isotropic states. In fact, we should worry only about boundedness on 
1733: those states which describe a collapsing universe which is homogeneous 
1734: and isotropic on large scales. Indeed, in this section we will 
1735: display the result of a calculation 
1736: which indicates that 
1737: the expectation value of the inverse scale factor with 
1738: respect to (kinematical) coherent states \cite{20} peaked on any 
1739: homogeneous and 
1740: isotropic point in phase space (e.g. a FRW universe along the classical 
1741: singular trajectory) {\bf is bounded from above in LQG even at the Big Bang.} 
1742: 
1743: The details of our calculation can be 
1744: read in the companion paper \cite{16}. However, again we can explain the 
1745: basic mechanism for this boundedness result in very intuitive terms:\\
1746: Spin network states are, roughly speaking, eigenstates for the kinematical
1747: geometric operators such as the triad operator out of which the inverse 
1748: scale factor is assembled. They are thus maximally sharp for half of the 
1749: degrees of freedom (the electric fields) but maximally unsharp for the 
1750: other half (the conections) which are encoded in the 
1751: holonomy operators. Coherent states intermediate in between those two 
1752: uncertainties and in fact minimize the Heisenberg uncertainty bound as far 
1753: as conjugate canonical pairs are concerned. This is achieved by 
1754: superimposing basis states such as spin network states with carefully 
1755: chosen coefficients. These coefficients depend on a point in the classical 
1756: phase space, in our case a field configuration $(A,E)$ of connections and 
1757: electric fields respectively. The coherent states are thus labelled by 
1758: these points and they are peaked in phase space in the sense that the 
1759: overlap function $|<\psi_{(A,E)},\psi_{(A',E')}>|^2$ vanishes 
1760: exponentially fast unless $||A-A'||^2,\;||E-E'||^2\le s$ where $||.||$ is 
1761: a suitable norm on the classical field space and $s$ is a parameter which 
1762: encodes the Gaussian width of the wave packet or the volume of a cell in 
1763: phase space (it is proportional to $\hbar$). The Gaussian decay of the 
1764: overlap function implies that in terms of spin network states the state is 
1765: of the form 
1766: $\psi_{(A,E)}=\sum_j\; c_j(A,E)\; T_j$ where we have not displayed the 
1767: graph label 
1768: for simplicity and where the $c_j$ decay (symbolically) as $\exp(-s j^2)$. 
1769: The label 
1770: $j$ is a compound label for all participating spins and intertwiners.
1771: Now while the norm of the inverse scale factor in the states $T_j$ is 
1772: unbounded from above, that bound only diverges with a finite power as 
1773: $j\to \infty$. This divergence therefore is completely swamped by the 
1774: Gaussian decay of the exponential. This holds for rather generic values
1775: of $A,E$ and in particular when we choose a homogeneous and isotropic 
1776: point in the phase space. Hence the mechanism is similar as in statistical
1777: physics: ``Energy'' (the argument of the Gaussian factor) wins over 
1778: ``entropy'' (the number of configurations leading to divergent 
1779: eigenvalues). This a rather different mechanism which does not require
1780: boundedness of the inverse scale factor.\\
1781: \\
1782: Let us go into more detail:\\  
1783: Given a point $(A_0,E_0)$ in the classical phase space, our (not 
1784: normalized) coherent states 
1785: are of the general form 
1786: \be \label{4.1}
1787: \psi_{(A_0,E_0)}=\sum_\gamma\;\psi_{\gamma,(A_0,E_0)}
1788: :=[e^{-\hat{{\rm\bf C}}/\hbar}\;\delta_{A'}]_{A'\mapsto A^\Cl(A_0,E_0)}
1789: \ee
1790: where the uncountable sum extends over all finite graphs and 
1791: $\psi_{\gamma,(A_0,E_0)}$ is a countably infinite coherent superposition 
1792: of 
1793: spin network states over $\gamma$. The coefficients in that superposition 
1794: are uniqely determined by the generator ${\rm\bf C}$ of the coherent 
1795: states, also 
1796: called the complexifier \cite{22a}, which is supposed to have the 
1797: dimension of an action. In fact, as we can see from 
1798: (\ref{4.1}), the coherent states are 
1799: nothing else than the heat kernel (with respect to the operator 
1800: $\hat{{\rm\bf C}}$) evolution of the $\delta$ distribution with support at 
1801: $A'$ 
1802: which is then analytically extended to $A^\Cl(A_0,E_0)$. The complexified 
1803: connection $A^\Cl$ also is generated by ${\rm\bf C}$ via
1804: \be \label{4.2}
1805: A^\Cl:=\sum_{n=0}^\infty \;\frac{i^n}{n!}\;\{{\rm\bf C},A\}_{(n)} 
1806: =A+i\;\{{\rm\bf C},A\}-\frac{1}{2}\;\{{\rm\bf C},\{{\rm\bf 
1807: C},A\}\}-\frac{i}{6} \{{\rm\bf C},\{{\rm\bf C},\{{\rm\bf C},A\}\}\}+...
1808: \ee
1809: This may look unfamiliar, however, this is precisely the way that the 
1810: unnormalized coherent states for the harmonic oscillator are constructed 
1811: for which the 
1812: complexifier reads ${\rm\bf C}=p^2/(2m\omega)$ and the classicality 
1813: parameter 
1814: is $s=\hbar/(m\omega)$ (the role of the connection is of 
1815: course played by the configuration coordinate $q$). 
1816: 
1817: Ideally, according to the subsequent section, ${\rm\bf C}$ should be 
1818: adapted 
1819: to a given Hamiltonian, but as we 
1820: said, here we will carry out a completely kinematical calculation to get a 
1821: first insight. It is therefore appropriate to choose ${\rm\bf C}$ in such 
1822: a way as 
1823: to obtain the simplest possible states. One such choice involves the 
1824: square of the area operator \cite{22a} which in turn is proportional to 
1825: $E^2$ and thus will give rise to states very similar to the ones for the 
1826: harmonic oscillator because in both cases the complexifiers are of the 
1827: momentum squared type giving rise to Gaussian wave packets. Rather than 
1828: defining ${\rm\bf C}$ we directly display 
1829: the corresponding states which in that case are, roughly, of the direct 
1830: product form
1831: \be \label{4.3}
1832: \psi_{\gamma,(A_0,E_0)}=\prod_{e\in E(\gamma)}\;\psi_{e,(A,E)}
1833: \mbox{ where } 
1834: \psi_{e,(A_0,E_0)}(A)=\sum_{2j=1}^\infty\;(2j+1)\;e^{-\frac{t(e)}{2}j(j+1)}
1835: \;\mbox{Tr}(\pi_j(g_e(A,E)\;A(e)^{-1}))
1836: \ee
1837: Here $\pi_j$ is the spin $j$ irreducible representation of $SU(2)$. 
1838: Notice that that the wave function $\psi_{e,(A_0,E_0)}(A)$ is a function 
1839: of $A$ (or of the $SU(2)$ valued holonomies $A(e)$) which depends 
1840: parametrically on $(A_0,E_0)$. The dimensionless parameter $t(e)$ depends 
1841: on the details of ${\rm\bf C}$ and satisfies $t(e)>0,\;t(e\circ 
1842: e')=t(e)+t(e'),\;
1843: t(e^{-1})=t(e)$. Finally, 
1844: $g_e(A_0,E_0)\approx \exp(i \tau_j E^j_0(S_e)/2)\;A_0(e)$ where 
1845: $S_e$ is, roughly speaking, the face of a cell complex dual to $\gamma$ 
1846: (i.e. each $S_e$ 
1847: intersects $\gamma$ only in an interior point of $e$) which also depends 
1848: on the details of ${\rm\bf C}$). 
1849: Experts will see the representation theory of $SU(2)$ and the Peter\&Weyl
1850: theorem at work in (\ref{4.3}). 
1851: 
1852: As explicitly shown in \cite{20}, expectation value computations with 
1853: coherent states are qualitatively unaffected when we replace the group
1854: $SU(2)$ by $U(1)^3$ for which however the analysis simplifies 
1855: tremendously. The corresponding coherent states are then given by
1856: \be \label{4.4}
1857: \psi_{e,(A_0,E_0)}(A)=\sum_{|n_1|,|n_2|,|n_3|=1}^\infty\;
1858: e^{-\frac{t(e)}{2}\sum_{j=1}^3 n_j^2}\;
1859: \;\prod_{j=1}^3\; (g^j_e(A,E)\;A^j(e)^{-1})^{n_j}
1860: \ee
1861: where $A^j(e)=\exp(i\int_e A^j)$ and  
1862: where 
1863: \be \label{4.5}
1864: g^j_e(A_0,E_0)=\exp(\int_{S_e} dS_a E^a_{0j}+i\int_e A_0^j)
1865: \ee
1866: To carry out concrete calculations we just have to insert concrete values 
1867: for $A_0,E_0$ into those formulae. In cosmological applications we choose,
1868: as in LQC (see section \ref{s0}), e.g. $A_{0a}^j=q\delta_a^j$ and 
1869: $E^a_{0j}=p\delta_a^j$ where 
1870: $a=\sqrt{|p|}$ is the scale factor. Thus in LQG, rather than performing a 
1871: {\it cancellation} of the inhomogeneous degrees of freedom as in LQC, we 
1872: perform a {\it suppression} of their fluctuations by using appropriate 
1873: states.
1874: 
1875: What one actually does in the computations is not to use
1876: the states $\psi_{(A_0,E_0)}$ but rather the ``cut -- off'' states
1877: $\psi_{\gamma,(A_0,E_0)}$ for finite but sufficiently fine $\gamma$.
1878: Roughly speaking, the analysis of \cite{20,22a} shows that the optimal 
1879: graph, for given resolution scale $L$ of a measurement\footnote{Given a 
1880: metric $g_0$ to be approximated, we want fluctuations in the length to be
1881: much smaller than $L$ where $L$ is to be measured with respect to $g_0$. On 
1882: the other hand, the Planck scale discreteness of LQG forces 
1883: $\epsilon\gg L$.} 
1884: should have an average edge length scale (with respect to the metric to be 
1885: approximated) of the order of a geometric mean $\epsilon\approx L^k 
1886: \ell_P^{1-k}$ where $0<k<1$ is a parameter that depends on the (partial)
1887: observables to be approximated. 
1888: The reason for working with the cut -- off states is that the 
1889: $\psi_{(A_0,E_0)}$ are not normalizable 
1890: on the kinematical Hilbert space due to the sum over the uncountable 
1891: number of graphs. They merely serve as a generator for the cut -- off 
1892: states which {\it are} normalizable.
1893: 
1894: What we have done then is to estimate from above the expectation value 
1895: not only of the inverse scale factor in LQG for a general valence $M$ of the vertex 
1896: for arbitrary $(A_0,E_0)$ but also the most general polynomials built out 
1897: of generalized co -- triad operators coming from classical expressions of 
1898: the form $(\{A,V^r\})^N$ which become upon quantization operators of the form 
1899: $\big[\frac{\mb{i}}{\hbar\kappa}\big]^N\prod_{I=1}^N\hat{q}^j_{e_I}(v,r)=
1900: \big[\frac{\mb{i}}{\hbar\kappa}\big]^N\prod_{I=1}^N\hat{h}^j_{e_I}\big[(\hat{h}^j_{e_I})^{-1},\hat{V}^r \big]$ where $r,N$ are fixed, positive, rational and natural numbers respectively which are determined by the concrete coupling between
1901: matter and gravity \cite{10b}. Here $\hat{h}^j_e$ are the holonomies of the connection $A$ along edges of the graph, the classical volume expression $V$ turns into the volume operator $\hat{V}$ and Poisson brackets into commutators. A bound is given by 
1902: \footnote{Neglecting all terms of order $t^\infty$.} 
1903: 
1904: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1905: % bound for general A_0, E_0
1906: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1907: 
1908: \ba\label{finale obere Schranke fuer Erwartungswert}
1909:     \lefteqn{\Big<~\Psi_{m,\gamma}^{(v)}(A)~\Big|~\prod_{k=1}^N 
1910: \hat{q}^{j_k}_{I_k} (r)~\Big|~\Psi_{m,\gamma}^{(v)}(A)~ \Big>\le}
1911:    \nonumber\\
1912:    &\le&
1913: %     \frac{
1914: (\ell_P)^{3rN}(9M)^N|Z|^\frac{rN}{2}\big[\frac{2\mb{A}}{T^2}\big]^N \;\;
1915: %}     {\prod_{I,i}[1+K_{t(I)}^{i}]}
1916:      \Bigg[\bigg[\frac{3Mp}{4T}\bigg]^N+\sum\limits_{n=1}^{N}\frac{N!}{(N-n)!~n!} \bigg[\frac{3Mp}{4T}\bigg]^{N-n}
1917:      \prod_{l=1}^n\bigg[\frac{3M+2(l-1)}{4}\bigg]\Bigg]~~~~~~~~~ 
1918: \ea
1919: 
1920: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1921: % explanation of terms
1922: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1923: 
1924: where $M$ denotes the valence of the vertex, $N$ the degree of the polynomial, $Z$ is a constant numerical prefactor, depending on the Immirzi parameter and the regularization of the volume operator. 
1925: 
1926: Moreover 
1927: \[\begin{array}{lllclc}
1928:                                 T:=\min\limits_I\{T_I\} &~~~~&\rightarrow~~
1929: 				&\mb{A}&:=1+\frac{p}{T}
1930: 				\\
1931: 				p:=\max\limits_{I,j}\{|p^j_I(m)|\}
1932: 				&&&          
1933:                                 
1934: 			     \end{array} \]
1935: 
1936:  with $T_I=\sqrt{t(e_I)}$, $t(e_I)$ is the classicality 
1937:  parameter mentioned above and
1938:  $p^j_I(m)\approx\frac{\epsilon}{L^3} \int_{S_{e_I}} dS_a E^a_j$.
1939:  Here $m=(A_0,E_0)=(A,E)$ is the point in classical phase space the 
1940:  coherent state is peaked on and $L,\;\epsilon$ respectively are the 
1941: resolution scale and graph scale mentioned above. 
1942: 
1943: Specializing to homogeneous data  
1944: $m=(A^j_{0a},E^a_{0j})=(\tilde{q}\delta^j_a,\tilde{p}\delta^a_j)$ gives 
1945:  $p=\tilde{p}\cdot\max\limits_{I}C_I^{(construct)}$, with numerical 
1946: constants
1947:  $C_I^{(construct)}$ carrying information about the details of the 
1948: coherent state. 
1949: 
1950: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1951: % bound for homogeneous data
1952: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1953: 
1954: Taking the limit $a\to 0$ which causes $p=0, \mb{A}=1$ demonstrates that nothing dramatic happens
1955: and we get the huge but finite bound 
1956: 
1957: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1958: % bound at a=0
1959: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1960: \ba\label{obere Schranke fuer Erwartungswert im Urknall}
1961:    \Big<~\Psi_{m,\gamma}^{(v)}(A)~\Big|~\prod_{k=1}^N \hat{q}^{j_k}_{I_k} (r)~\Big|~\Psi_{m,\gamma}^{(v)}(A)~ \Big>\bigg|_{a=0}   
1962:    &\le&
1963: %     \frac{
1964: (\ell_P)^{3rN}(9M)^N|Z|^\frac{rN}{2}\big[\frac{2}{T^2}\big]^N \;\;
1965: %}{\prod_{I,i}[1+K_{t(I)}^{i}]}
1966:       \prod_{l=1}^N\bigg[\frac{3M+2(l-1)}{4}\bigg]~~~~~~~~~ 
1967: \ea
1968: 
1969: 
1970:  For the case of the inverse scale factor we have to set $(r=1/2,\;N=6)$ 
1971: and to multiply by $\frac{1}{(\hbar\kappa)^N}=\frac{1}{(\ell_P)^{12}}$. \\
1972: \\
1973: These results prove our claim: The expectation value (and thus also the 
1974: norm, just double $N$) of the inverse scale factor is bounded with respect 
1975: to kinematical semiclassical states even when the classical metric that 
1976: labels the coherent state becomes maximally degenerate. As in LQC this is 
1977: a strong indication 
1978: for a drastic modification of the effective Friedmann equations near 
1979: $a=0$ \cite{13}. This should be worked out in more detail in order to 
1980: decide, along the scheme given in section \ref{s3}, whether the initial 
1981: local singularity is avoided or not.\\
1982: \\
1983: Remarks:\\ 
1984: 1.\\
1985: We have computed the expectation value of the 
1986: analogon of the inverse 
1987: scale factor with respect to kinematical coherent states whose peak 
1988: follows the classcial (singular) trajectory. One should do the same with 
1989: respect to the matter degrees of freedom. However, it is clear from the 
1990: outset that this will not change our result: The classical Friedmann 
1991: equations in the gauge outlined above are 
1992: \be \label{4.6}
1993: 3H^2=8\pi G(\frac{\pi^2}{2a^6}+\frac{a^3 V}{a^3})+\Lambda-\frac{3k}{a^2},\;\;
1994: \dot{\pi}=-a^3 V'
1995: \ee
1996: where $H=\frac{\dot{a}}{a}$ is the Hubble function and $\pi=a^3 
1997: \dot{\phi}$
1998: the conjugate momentum. Consider the case of a 
1999: potential which depends polynomially on $\phi$ but not exponentially.
2000: We want to solve (\ref{4.6}) close to $a=0$. We now make the 
2001: self-consistent assumption (to be checked) that $a^3 V$ (and thus $a^3 
2002: V'$) 
2003: vanishes in the 
2004: limit $a\to 0$. Hence $\pi$ becomes constant close to $a=0$ according to 
2005: the 
2006: second equation in (\ref{4.6}). Then we may neglect the 
2007: potential ($V$), cosmological 
2008: ($\Lambda$) and curvature ($k$) term in the first equation as $a\to 0$
2009: and thus we may solve $a(t)\propto t^{1/3}$ and $\phi(t)\propto \ln(t)$.
2010: Hence indeed $a^3 \phi^n\propto t [\ln(t)]^n$ vanishes at $t=0$ and our 
2011: assumption was correct. It follows that the expectation value with respect 
2012: to coherent states, taking into account both matter and geometry, of the 
2013: scalar contribution to the Hamiltonian constraint 
2014: (\ref{0.7}), whose homogeneous reduction is 
2015: $\pi^2/(2a^3)+a^3 V$, is indeed non-singular even when the coherent states 
2016: are evaluated along the {\it exact} classical trajectory, if and only if 
2017: the expectation value of the analogon in (\ref{3.3}) of $1/a^3$ is bounded 
2018: from 
2019: above. This is precisly what we verified. \\
2020: 2.\\
2021: We have shown unboundedness of the inverse scale factor in LQG (with gauge 
2022: group $SU(2)$) and did expectation value calculations therefof for an 
2023: approximation to LQG using $U(1)^3$ instead of $SU(2)$. It is maybe 
2024: worthwhile to mention that the inverse scale factor for that approximation 
2025: is also unbounded on zero volume eigenstates although one has to use 
2026: graphs of valence at least five here which is due to the detailed 
2027: difference between the non -- Abelean group $SU(2)$ and the Abelean group
2028: $U(1)^3$.\\
2029: 3.\\
2030: Notice that the bound in formula (\ref{finale obere Schranke fuer 
2031: Erwartungswert}) is completely general. It does not depend on the value of 
2032: the classical connection $A$ that labels the coherent state, it only 
2033: depends on the electric field $E$. Setting $E^a_j=0$ as appropriate 
2034: for the isotropic singularity gives the  
2035: upper bound (\ref{obere Schranke fuer Erwartungswert im Urknall}). 
2036: However, we may consider any other homogeneous model, for instance 
2037: inhomogeneous ones such as Bianchi I (Kasner), Bianchi IX or Mixmaster 
2038: Models (BKL scenario). In these cases the different components of the 
2039: three metric vanish with different rates as we approch the singularity
2040: (or do not vanish at all or might even diverge, at least in the absence 
2041: of matter). As we can see from the above general estimate, the important 
2042: question is whether the quantity 
2043: $p$ defined above, which is proportional to $E^a_j=\sqrt{\det(q)} e^a_j$,
2044: remains finite in all those cases. For instance, in the Kasner 
2045: models with 
2046: line element $ds^2=-dt^2+\sum_{a=1}^3 t^{2p_a} (dx^a)^2$ where the 
2047: real parameters $p_a$ are subject to $\sum_a p_a=\sum_a p_a^2=1$ we find 
2048: in the diagonal gauge $E^a_j=\delta^a_j t^{1-p_a}$ which stays bounded 
2049: for $t\to 0$ because $p_a\le 1$. In particular, contracting 
2050: directions (with increasing $t$) are allowed. This covers all 
2051: Kasner models and shows that the expectation value calculation 
2052: performed here also hints at local singularity 
2053: avoidance for at least some class of anisotropic models. 
2054: Whether this holds for all homogeneous Bianchi models and also after
2055: matter is coupled is a task left to the reader.
2056: 
2057: 
2058: 
2059: 
2060: 
2061: \section{Scheme for Determining the Presence or Absence of the Initial 
2062: Singularity}
2063: \label{s3}
2064: 
2065: The computation of the previous section is very encouraging.
2066: Unfortunately, it does not yet prove that in LQG the 
2067: Big Bang singularity or any other classical spacetime curvature 
2068: singularity is 
2069: quantum mechanically absent. As we have explained at length in the 
2070: introduction, fundamentally, singularity avoidance 
2071: must be formulated in terms of physical states\footnote{The singularity 
2072: effects A and B for the global singularity have indeed been discussed for 
2073: solutions of the Wheeler DeWitt equation in LQC. However, that does not 
2074: grant
2075: that a sufficient number of them are normalizable or have non vanishing 
2076: norm in the physical inner product. Once again, formal solutions are not 
2077: automatically physical states.}
2078: and physical operators 
2079: which has not been done so far even in LQC\footnote{In terms of the 
2080: the partial and complete observables discussed below, there is in general
2081: no relation between the spectra of the complete and partial observables
2082: \cite{20a}. Hence kinematical boundedness does not imply physical
2083: boundedness as far as the local singularity is concerned.}
2084: because the volume, the inverse scale factor and the value of the inflaton 
2085: field etc. are not gauge invariant quantities (they do not commute with 
2086: e.g. the Hamiltonian constraint). 
2087: 
2088: Fundamentally, what one would need to do 
2089: in order to establish a singularity avoidance result is something along 
2090: the following lines\footnote{This is just a sketch of ideas, not a 
2091: ``paradigm''.}: 
2092: %
2093: \begin{itemize}
2094: %
2095: \item[1.] Ideally one should have sufficiently explicit knowledge about 
2096: the 
2097: physical Hilbert space ${\cal H}_{Phys}$ and physical states. An avenue 
2098: towards this goal is the Master Constraint Programme \cite{11}. Notice 
2099: that physical states automatically solve the Quantum Einstein Equations.
2100: That means that the quantum gauge evolution \cite{10b}, defined 
2101: similarly as in \cite{12} as the coefficients of the physical states 
2102: when expanded into (spatially diffeomorphism invariant) spin network 
2103: functions, is non -- singular by definition thanks to the fact that 
2104: operators
2105: like the inverse scale factor are densely defined in LQG. 
2106: Moreover, in LQG there are 
2107: physical states which are composed of spin network states 
2108: that describe a sign flip with respect to the expectation value of the 
2109: triad orientation sign operator \cite{20b} and thus may describe quantum 
2110: geometries which include a Pre Big Bang in the sense of \cite{12}.
2111: As we have already outlined in the introduction, the global 
2112: singularity is avoided if and only if there are sufficiently many 
2113: semiclassical of these physical states. 
2114: %
2115: \item[2.] (In)famously, canonical quantum gravity, especially in the
2116: cosmological context, is a theory without a canonical Hamiltonian. 
2117: However, one can construct physical Hamiltonians using the relational 
2118: point of view \cite{15,15a}. These are Dirac observables by construction.
2119: They are obtained by i) using non -- gauge invariant quantities $P,T$
2120: where $P$ is called a partial observable and $T$ a collection 
2121: of\footnote{In general one needs as many clock variables as there are 
2122: constraints.} clock variables,
2123: ii) computing the gauge flows (Hamiltonian evolution with respect 
2124: to the Hamiltonian constraints\footnote{There are also as many unphysical 
2125: time parameters $t$ as there are constraints. In GR they become 
2126: functions and are called lapse and shift.}) 
2127: $t\mapsto 
2128: \alpha_t(P),\;\;\alpha_t(T)$, 
2129: iii) choosing parameters $\tau$ and solving the equations 
2130: $\alpha_t(T)=\tau$ for $t$ and iv) inserting those values of $t$ 
2131: into 
2132: $\alpha_t(P)$. The resulting object $P_T(\tau)$ has the physical 
2133: interpretation of giving the value of $P$ at the moment (gauge parameter 
2134: $t$) when $T$ has attained the value $\tau$. The function 
2135: $P_T(\tau)$ is called the corresponding complete
2136: or Dirac observable. The physical Hamiltonians $H_T(\tau)$ are in general
2137: explicitly dependent on the physical time $\tau$ and are defined by 
2138: $\{H_T(\tau),P_T(\tau)\}:=d/d\tau P_T(\tau)$ for all $P$. Of course there 
2139: are many 
2140: possible choices for $H_T$ depending on the choice of the ``clocks'' $T$
2141: (in GR we need an infinite number of those). In cosmology it would be 
2142: nice to use the scale factor (total volume of space) as a clock. 
2143: If a corresponding positive (or at least semibounded) and physical time 
2144: independent Hamiltonian exists and can be 
2145: defined as a self -- adjoint operator on the physical Hilbert space then
2146: we may use it to define evolution on the physical Hilbert space. 
2147: %
2148: \item[3.] If a physical Hamitonian is given, then we will need coherent 
2149: states in ${\cal H}_{Phys}$ which are preserved by $\hat{H}_T$ for 
2150: sufficiently long physical time. A general 
2151: idea for how to do that has been given by Klauder \cite{20} at least for 
2152: systems with a finite number of degrees of freedom.
2153: %
2154: \item[4.] We are now in the position to address the issue about the 
2155: local singularity: Given a physical coherent state $\psi$ peaked, on 
2156: scales 
2157: large 
2158: compared to the Planck scale, corresponding to $\tau=\tau_0\gg \tau_P$, on 
2159: classical 
2160: singular initial data, consider 
2161: the map 
2162: $\tau\mapsto <\psi,\widehat{P_T(\tau-\tau_0)}\psi>_{Phys}=
2163: <\psi(\tau-\tau_0),\widehat{P_T(\tau_0)}\psi(\tau-\tau_0)>_{Phys}$
2164: with $\psi(\tau):=e^{i\tau\hat{H}_T}\psi$. Let $P$ be any classically 
2165: singular quantity, say the 
2166: matter energy density or a curvature scalar density (suitably smeared). 
2167: Then the local 
2168: singularity, with respect to $P$, is absent provided that
2169: $<\psi,\widehat{P_T(-\tau_0)}\psi>_{Phys}$ is finite (assuming that 
2170: $\tau=0$ corresponds to vanishing classical volume which will be the 
2171: case if we choose $V$ as one of the clock variables). 
2172: %
2173: \end{itemize} 
2174: %
2175: Notice that this ambitious scheme has not been followed even in the 
2176: simplified context of LQC where, however, all the steps could presumably 
2177: be carried out sufficiently explicitly, see \cite{22b} for first steps 
2178: into that direction. In full LQG these steps cannot be 
2179: carried out exactly because the theory is too complicated. Thus, in 
2180: order to make progress one must use suitable approximation schemes. A 
2181: possible starting point could be the following:
2182: %
2183: \begin{itemize}
2184: %
2185: \item[1'.] Using kinematical coherent states $\psi_{A,E}$ peaked on the 
2186: constraint 
2187: surface of the phase space we arrive at a subspace of the kinematical 
2188: Hilbert space ${\cal H}_{Kin}$ with the property that the constraints 
2189: are satisfied approximately in the sense of expectation values, that is
2190: $<\psi,\hat{C}\psi>\approx 0$. Denote that subspace 
2191: by ${\cal H}'_{Phys}$ \footnote{Alternatively, find an approximation 
2192: method for the Master Constraint 
2193: Programme (also using kinematical coherent states and path integral 
2194: methods, see \cite{11}).}. In systems where one can compare predictions by 
2195: the exact ${\cal H}_{Phys}$ with those by ${\cal H}'_{Phys}$ it is often 
2196: the case that the qualitative answers do not differ much. See e.g. 
2197: \cite{22}. Since there are as many semiclassical kinematical states as
2198: classical initial data (classical spacetimes), global singularities cannot 
2199: be 
2200: determined within this approximate scheme which uses the kinematical 
2201: Hilbert space.
2202: %
2203: \item[2'.] The formulas for $P_T(\tau)$ and $H_T$ are infinite power 
2204: expansions in the ``parameters'' $\tau-T$ \cite{15a}. Hence, close to the 
2205: singularity,
2206: i.e. when $\tau\approx T=0$ (this has to be understood ultimatively in 
2207: terms 
2208: of operators; we assume that the $T,\tau$ are chosen such that 
2209: $T=\tau=0$ corresponds to the singularity) we may 
2210: approximate those formulas by the first few terms and 
2211: get approximate expressions $\hat{P}'_T(\tau),\;\hat{H}'_T$ that one can 
2212: hope to handle. These approximations should be appropriate in order to 
2213: answer qualitatively questions about the very early universe.
2214: %
2215: \item[3'.] The coherent states of item 1' should ideally be chosen as to 
2216: be preserved by the approximate Hamiltonian determined in step 2'. If that 
2217: is too difficult, peak the kinematical coherent states of step 1' on
2218: classically singular initial data and then quantum evolve them with the 
2219: (approximate) Hamiltonian $\hat{H}'_T$, that is 
2220: $\tau\mapsto \exp(i\tau \hat{H}'_T)\psi_{A,E}$.
2221: %
2222: \item[4'.] This step is identical to step 4 above with the obvious 
2223: replacements. 
2224: %
2225: \end{itemize}
2226: %
2227: Notice that one could consider an even easier scheme: Just take any set of 
2228: kinematical coherent states not generated by any (approximate) physical 
2229: Hamiltonian and consider the unphysical time evolution 
2230: $t\mapsto \psi_{A_0,E_0}(t):=\psi_{A(t),E(t)}$ where $A(t),E(t)$ is 
2231: a one 
2232: parameter classical gauge 
2233: evolution of the initial data under the flow of the 
2234: corresponding Hamiltonian 
2235: constraint (not of any physical Hamiltonian)\footnote{This 
2236: corresponds to a specific choice of lapse and shift.} with initial
2237: condition $A(t_0)=A_0,\;E(t_0)=E_0$ and $t_0\gg t_P$. 
2238: Then take any quantity $P$ 
2239: which is not necessarily a (approximate) Dirac observable and study
2240: $<\psi_{A(t),E(t)},\hat{P}\psi_{A(t),E(t)}>_{Kin}$ in the limit $t\to 0$
2241: corresponding to the classical singularity (in our case $A(t)$ diverges 
2242: while $E(t)$ vanishes at $t=0$). This is essentially what we 
2243: did in the previous section where $P$ played the role of the 
2244: inverse scale factor and the coherent states we used are those of 
2245: \cite{20}. Wile in this case we obtained a boundedness result, one may not 
2246: draw any conclusions from that as the following example reveals:\\
2247: Consider the Hamiltonian of the hydrogen atom. Let 
2248: $\psi_{\vec{q},\vec{p}}$ be the coherent states of the harmonic oscillator 
2249: peaked at the point $(\vec{q},\vec{p})$ in phase space. Let 
2250: $\vec{q}(t),\vec{p}(t)$ be the trajectory of a classical electron falling
2251: radially into the proton from a distance large compared to the atomic 
2252: radius where $t=0$ is the point of time when the 
2253: electron reaches the center. Then it is easy to see that 
2254: $<\psi_{\vec{q}(t),\vec{p}(t)},\hat{H} \psi_{\vec{q}(t),\vec{p}(t)}>$ 
2255: diverges as $t\to 0$. This is because the electron becomes infinitely fast 
2256: at the center and the expectation value of the kinetic term diverges
2257: (while the expectation value of the potential energy stays finite). In 
2258: contrast, the expectation 
2259: value
2260: $<\psi_{\vec{q},\vec{p}}(t),\hat{H} \psi_{\vec{q},\vec{p}}(t)>
2261: =<\psi_{\vec{q},\vec{p}},\hat{H} \psi_{\vec{q},\vec{p}}>$ is finite where 
2262: $\psi_{\vec{q},\vec{p}}(t):=\exp(it\hat{H})\psi_{\vec{q},\vec{p}}$
2263: is the quantum evolved state\footnote{Remarkably, the states 
2264: $\psi_{\vec{q},\vec{p}}(t)$ and $\psi_{\vec{q}(t),\vec{p}(t)}$ precisely 
2265: coincide if and only if the underlying dynamics is that of a finite or 
2266: infinite sum of uncoupled harmonic oscillators.}. Thus, while 
2267: $\psi_{\vec{q}(t),\vec{p}(t)}$ is nicely peaked on the classical 
2268: trajectory, it does not correspond to the true quantum mechanical time 
2269: evolution. On the other hand, the true quantum mechanical time evolution 
2270: $\psi_{\vec{q},\vec{p}}(t)$ usually does not stay peaked on the classical 
2271: trajectory for sufficiently long time in order to qualify as a 
2272: semiclassical state representing the classical evolution far away from 
2273: the deep quantum regime closely enough, unless the coherent state is 
2274: actually adapted to the Hamiltonian $\hat{H}$. \\
2275: \\
2276: Summarizing, ideally
2277: the coherent states must be chosen according to the Hamiltonian in 
2278: question in order to arrive at reliable predictions. If that is too hard, 
2279: then at least the kinematical coherent state should be evolved quantum 
2280: mechanically but only as long as it remains close to the classical 
2281: trajectory for scales above which the classical theory is reliable. In 
2282: contrast, the classically evolved 
2283: kinematical semiclassical state not constructed according to the 
2284: Hamiltonian is inappropriate in general which is why the calculation of 
2285: section \ref{s4} is promising but inconclusive.
2286: 
2287: 
2288: 
2289: \section{Conclusions}
2290: \label{s5}
2291: 
2292: 
2293: Concluding, the issue of the presence or absence of the initial
2294: singularity in full LQG remains open. The results of LQC are 
2295: important, they are very 
2296: promising, provide important consistency checks for LQG and produce new 
2297: insights and ideas about LQG. However, as we 
2298: hope to have demonstrated, they are so far inconclusive for LQG. 
2299: Therefore sentences such as ``LQC is the cosmological 
2300: sector of LQG'' or ``in LQG the Big Bang is absent'' \cite{23}
2301: are misleading for the non expert and should be used with due care.
2302: 
2303: As our detailed analysis 
2304: reveals, one must pay attention to the QFT aspects of the full theory 
2305: and cannot rely on the analysis of a quantum mechanical toy model.
2306: This is, unfortunately, different from the case of the hydrogen atom 
2307: where experiments confirmed from the outset that the quantum mechanical 
2308: toy model was close to the truth and that the more accurate QED should 
2309: only give corrections. 
2310: It may still be that LQC is qualitatively right but if that 
2311: is the case then the mechanism for singularity resolution is very 
2312: different in full LQG and much more elaborate, if LQG is a valid candidate 
2313: for quantum gravity at all. 
2314: 
2315: In order to resolve the issue about the initial singularity and other 
2316: cosmological questions in full LQG we have proposed an ideal and 
2317: approximate programme, highlighting the importance of having control over 
2318: the physical Hilbert space and the physical observables. In a zeroth order 
2319: step of that programme we found 
2320: hints pointing at a resolution of the local initial singularity in LQG by 
2321: a very 
2322: different mechanism. But it is necessary to carry out all the steps before 
2323: definite conclusions can be reached.
2324: Since this programme was not even completed in LQC, again 
2325: the LQC model could be a good starting point to see where the ideal path
2326: leads us in a solvable situation. As far as the full theory is 
2327: concerned we will only be able to follow the approximate path and we must 
2328: learn how to control the corresponding approximation errors. \\
2329: \\
2330: In summary, the investigation of the comological or black hole sector of 
2331: LQG, about 
2332: which the LQG inspired LQC model has taught us important first lessons,  
2333: is an ambitious and  
2334: fascinating research field which will require the joint effort of 
2335: both cosmologists and quantum geometers in the close future. \\
2336: \\ \\ 
2337: \\
2338: {\large Acknowledgements}\\
2339: \\
2340: It is our pleasure to thank Martin Bojowald for many clarifying comments 
2341: and a careful reading of the manuscript.
2342: J.B. thanks the Gottlieb Daimler-- and Karl Benz--Foundation and the DAAD 
2343: (German Academic 
2344: Exchange Service) for financial support. This work was supported in part
2345: by a grant from NSERC of Canada to the Perimeter Institute for Theoretical 
2346: Physics.
2347: 
2348: 
2349:  
2350:  
2351: 
2352: 
2353: 
2354: 
2355: 
2356: \begin{thebibliography}{99}
2357: 
2358: \parskip -5pt
2359: 
2360: \bibitem{1} C. Rovelli, ``Quantum Gravity'', Cambridge University
2361: Press, Cambridge, 2004\\
2362: T. Thiemann, ``Modern Canonical Quantum General Relativity'', Cambridge 
2363: University Press, 2005; gr-qc/0110034
2364: 
2365: \bibitem{1a} 
2366: C. Rovelli, ``Loop Quantum Gravity", Living Rev. Rel. {\bf 1} (1998) 1,
2367: gr-qc/9710008\\
2368: T. Thiemann,``Lectures on Loop Quantum Gravity'', Lecture Notes in
2369: Physics, {\bf 631} (2003) 41 -- 135, gr-qc/0210094\\
2370: A. Ashtekar, J. Lewandowski, ``Background Independent Quantum Gravity:
2371: A Status Report'', Class. Quant. Grav. {\bf 21} (2004) R53;
2372: [gr-qc/0404018]\\
2373: L. Smolin, ``An Invitation to Loop Quantum Gravity'', hep-th/0408048
2374:                                                                                 
2375: \bibitem{2} 
2376: R. Brunetti, K. Fredenhagen,, R. Verch, ``The Generally Covariant Locality
2377: Principle: A New Paradigm for Local Quantum Field Theory'',
2378: Commun. Math. Phys. {\bf 237}, 31-68,2003, [math-ph/0112041]
2379: 
2380: \bibitem{3} O. Lauscher, M. Reuter, ``Towards Nonperturbative
2381: Renormalizability of Quantum Einstein Gravity",
2382: Int. J. Mod. Phys. {\bf A17} 993 (2002), [hep-th/0112089];
2383: ``Is Quantum Gravity Nonperturbatively Renormalizable ?",
2384: Class. Quant. Grav. {\bf 19} 483 (2002), [hep-th/0110021]
2385: 
2386: \bibitem{4} 
2387: M. Bojowald, H. Morales-Tecotl, ``Cosmological Applications of Loop
2388: Quantum Gravity'', Lect. Notes Phys. {\bf 646} (2004) 421-462;
2389: [gr-qc/0306008]
2390: 
2391: \bibitem{5} 
2392: J. Halliwell , ``Derivation of the Wheeler-De Witt Equation from a Path 
2393: Integral for Minisuperspace Models'', Phys.Rev. {\bf D38} (1988) 2468 
2394: 
2395: \bibitem{5b} 
2396:  H. Sahlmann, ``When do Measures on the Space of Connections
2397: Support the Triad Operators of Loop Quantum Gravity?'', gr-qc/0207112;
2398: ``Some Comments on the Representation Theory of the Algebra Underlying
2399: Loop Quantum Gravity'', gr-qc/0207111\\
2400: H. Sahlmann, T. Thiemann, ``On the Superselection Theory of
2401: the Weyl Algebra for Diffeomorphism Invariant Quantum Gauge Theories'',
2402: gr-qc/0302090;
2403: ``Irreducibility of the Ashtekar-Isham-Lewandowski Representation'',
2404: gr-qc/0303074\\
2405: A. Okolow, J. Lewandowski, ``Diffeomorphism Covariant
2406: Representations of the Holonomy Flux Algebra'', gr-qc/0302059\\
2407: C. Fleichhack, ``Representations of the Weyl Algebra in Quantum
2408: Geometry'', math-ph/0407006\\
2409: J. Lewandowski, A. Okolow, H. Sahlmann, T. Thiemann, ``Uniqueness of
2410: Diffeomorphism Invariant States on Holonomy -- Flux
2411: Algebras'', to appear
2412:                                                                                 
2413: \bibitem{5a} 
2414: A. Ashtekar, C.J. Isham, ``Representations of the Holonomy
2415: Algebras of Gravity and Non-Abelean Gauge Theories",
2416: Class. Quantum Grav. {\bf 9} (1992) 1433, [hep-th/9202053]\\
2417: A. Ashtekar, J. Lewandowski, ``Representation
2418: theory of analytic Holonomy $C^\star$ algebras", in ``Knots and
2419: Quantum Gravity", J. Baez (ed.), Oxford University Press, Oxford 1994
2420: 
2421: \bibitem{6} W. Thirring, ``Lehrbuch der Mathematischen Physik", vol. 3,
2422: Springer Verlag, Berlin, 1978
2423: 
2424: \bibitem{7} F. Acerbi, G. Morchio, F. Strocchi, ``Infrared Singular Fields 
2425: and Nonregular Representations of CCR Algebras'', J. Math. Phys. {\bf 34}
2426: (1993) 899 -- 914
2427: 
2428: \bibitem{8} W. Thirring, H. Narnhofer, ``Covariant QED without indefinite 
2429: Metric'', Rev. Math. Phys. {\bf SI1} (1992) 197 -- 211
2430: 
2431: \bibitem{9} T. Thiemann, ``The LQG String: Loop Quantum Gravity 
2432: Quantization of String Theory I. Flat Target Space'', hep-th/0401172
2433: 
2434: \bibitem{10} 
2435:  T. Thiemann, ``Anomaly-free Formulation of non-perturbative,
2436: four-dimensional Lorentzian Quantum Gravity", Physics Letters {\bf B380}
2437: (1996) 257-264, [gr-qc/9606088]\\
2438: T. Thiemann, ``Quantum Spin Dynamics (QSD)",
2439: Class. Quantum Grav. {\bf 15} (1998) 839-73, [gr-qc/9606089]
2440: 
2441: \bibitem{10a} 
2442: T. Thiemann, ``Quantum Spin Dynamics II. The Kernel of the Wheeler-DeWitt 
2443: Constraint Operator",
2444: Class. Quantum Grav. {\bf 15} (1998) 875-905, [gr-qc/9606090]
2445: 
2446: \bibitem{10b} T. Thiemann, ``Quantum Spin Dynamics
2447: V. Quantum Gravity as the Natural Regulator of the Hamiltonian
2448: Constraint
2449: of Matter Quantum Field Theories",
2450: Class. Quantum Grav. {\bf 15} (1998) 1281-1314, [gr-qc/9705019]
2451: 
2452: \bibitem{10c} T. Thiemann, ``Quantum Spin Dynamics 
2453: IV. 2+1 Euclidean Quantum Gravity as a model to test 3+1
2454: Lorentzian Quantum Gravity", Class. Quantum Grav. {\bf 15} (1998)
2455: 1249-1280, [gr-qc/9705018]
2456: 
2457: 
2458: \bibitem{11} 
2459:  T. Thiemann, ``The Phoenix Project: Master Constraint
2460: Programme for Loop Quantum Gravity'', gr-qc/0305080\\
2461: B. Dittrich, T. Thiemann, ``Testing the Master Constraint Programme for 
2462: Loop Quantum Gravity 
2463: I. General Framework'', gr-qc/0411138;
2464: ``II. Finite Dimensional Systems'', gr-qc/0411139;
2465: ``III. SL(2,R) Models'', gr-qc/0411140;
2466: ``IV. Free Field Theories'', gr-qc/0411141;
2467: ``V. Interacting Field Theories'', gr-qc/0411142
2468: 
2469: \bibitem{11a} T. Thiemann, ``A Length Operator for Canonical Quantum 
2470: Gravity'', J. Math. Phys. {\bf 39} (1998) 3372-3392; [gr-qc/9606092]
2471: 
2472: \bibitem{12} M. Bojowald, ``Absence of Singularity in Loop Quantum 
2473: Cosmology'', Phys. Rev. Lett. {\bf 86} (2001) 5227-5230; [gr-qc/0102069]
2474: 
2475: \bibitem{13a} M. Bojowald, ``Homogeneous Loop Quantum Cosmology'',
2476: Class. Quant. Grav. {\bf 20} (2003) 2595; [gr-qc/0303073]
2477: 
2478: \bibitem{17a}  R. M. Wald, ``General Relativity", The University of 
2479: Chicago Press, Chicago, 1989
2480: 
2481: \bibitem{13} M. Bojowald, ``Inflation from Quantum Geometry'',
2482: Phys. Rev. Lett. {\bf 89} (2002) 261301, [gr-qc/0206054]
2483: 
2484: \bibitem{Bojowald1} M. Bojowald, ``Isotropic Loop Quantum Cosmology'',
2485: Class. Quant. Grav. {\bf 19} (2002) 2717; [gr-qc/0202077]
2486: 
2487: \bibitem{14} 
2488: V. Husain, O. Winkler, ``On Singularity Resolution in Quantum Gravity'' 
2489: Phys. Rev. {\bf D69} (2004) 084016; [gr-qc/0312094]\\
2490: ``Quantum Resolution of Black Hole Singularities'', [gr-qc/0410125]\\
2491: S. Hofmann, O. Winkler, ``The Spectrum of Fluctuations in Inflationary 
2492: Quantum Cosmology'', astro-ph/0411124
2493: 
2494: \bibitem{14a}
2495: K.V. Kuchar, M.P. Ryan, ``Is Minisuperspace Quantization Valid? Taub in 
2496: Mixmaster'', Phys. Rev. {\bf D40} (1989) 3982-3996\\
2497: S. Sinha, B.L. Hu, ``Validity of the Minisuperspace Approximation: An 
2498: Example from Interacting Quantum Field 
2499: Theory'', Phys. Rev. {\bf D44} (1991) 1028-1037\\
2500: A. Ishikawa, T. Isse, ``The Stability of Minisuperspace'' 
2501: Mod. Phys. Lett. {\bf A8} (1993) 3413-3428; [gr-qc/9308004]\\
2502: O. Saremi, ``Is Minisuperspace Quantum Gravity Reliable?'', gr-qc/0101108
2503: 
2504: \bibitem{14b} 
2505: D. Green, W. Unruh, ``Difficulties With Closed Isotropic Loop Quantum 
2506: Cosmology'', Phys. Rev. {\bf D70} (2004) 103502; [gr-qc/0408074]\\
2507: D. Cartin, G. Khanna, ``Absence of Pre-Classical Solutions in Bianchi I
2508: Loop Quantum Cosmology'', [gr-qc/0501016]
2509: 
2510: \bibitem{Bojowald} M. Bojowald, private communication
2511: 
2512: \bibitem{16} J. Brunnemann, T. Thiemann, ``Unboundedness of Triad -- Like 
2513: Operators in Loop Quantum Gravity'', gr-qc/0505033
2514: 
2515: \bibitem{15} C. Rovelli, in ``Conceptual Problems in Quantum Gravity'', 
2516: ed. by A. Ashtekar and J. Stachel, Birh\"auser, Boston, 1991, p. 126;
2517: ``What is Observable in Classical and Quantum Gravity?'', Class. Quantum 
2518: Grav. {\bf 8} (1991) 1895;
2519: ``Partial Observables'', Phys. Rev. {\bf D65} (2002) 124013
2520: 
2521: \bibitem{15a}
2522: B. Dittrich, ``Partial and Complete Observables for
2523: Hamiltonian Constrained Systems'', gr-qc/0411013\\
2524: T. Thiemann, ``Reduced Phase Space Quantization and Dirac Observables'',
2525: gr-qc/0411031
2526: 
2527: \bibitem{20a} B. Dittrich, T. Thiemann, ``Facts and Fiction about Dirac 
2528: Observables in (Loop) Quantum Gravity'', to appear
2529: 
2530: \bibitem{20d} 
2531: L. Modesto, ``Disappearance of Black Hole Singularity in Quantum 
2532: Gravity'', Phys. Rev. {\bf D70} (2004) 124009, [gr-qc/0407097];
2533: ``Kantowski-Sachs Space-Time in Loop Quantum Gravity'',
2534: gr-qc/0411032; ``Quantum Gravitational Collapse'', gr-qc/0504043\\
2535: A. Ashtekar, M. Bojowald, ``Black Hole Evaporation: A Paradigm'',
2536: gr-qc/0504029
2537: 
2538: \bibitem{20c}
2539: V. Husain, O. Winkler, ``On Singularity Resolution in Quantum Gravity''
2540: Phys. Rev. {\bf D69} (2004) 084016, [gr-qc/0312094];
2541: ``Quantum Resolution of Black Hole Singularities'', gr-qc/0410125;
2542: ``Quantum Black Holes'', gr-qc/0412039
2543: 
2544: \bibitem{16a}  H. Sahlmann, T. Thiemann, ``Towards the QFT on
2545: Curved Spacetime Limit of QGR. 1. A General Scheme'', [gr-qc/0207030];
2546: ``2. A Concrete Implementation", [gr-qc/0207031]
2547: 
2548: \bibitem{17} T. Thiemann, ``Closed Formula for the Matrix Elements of the 
2549: Volume Operator in Canonical Quantum Gravity'',
2550: J. Math. Phys. {\bf 39} (1998) 3347-3371; gr-qc/9606091\\
2551: J. Brunnemann, T. Thiemann, ``Simplification of the Spectral Analysis 
2552: of the Volume Operator in Loop Quantum Gravity'', gr-qc/0405060
2553: 
2554: \bibitem{de Pietri}R. De Pietri, {\it Spin Networks and Recoupling in Loop Quantum Gravity}, 
2555: [arXiv:~gr-qc/9701041]
2556: 
2557: \bibitem{17ba} A. Ashtekar, M. Bojowald, J. Lewandowski, ``Mathematical 
2558: Structure of Loop
2559: Quantum Cosmology'', Adv. Theor. Math. Phys. {\bf 7} (2003)
2560: 233,[gr-qc/0304074]
2561: 
2562: \bibitem{17c}
2563: M. Gaul, C. Rovelli, ``A Generalized Hamiltonian Contraint
2564: Operator in Loop Quantum Gravity and its Simplest Euclidean Matrix 
2565: Elements",
2566: Class. Quant. Grav. {\bf 18} (2001) 1593-1624; [gr-qc/0011106]
2567: 
2568: \bibitem{17d} 
2569: M. Bojowald, H. Kastrup, ``Quantum Symmetry Reduction for
2570: Diffeomorphism Invariant Theories of Connections", Class. Quantum Grav.
2571: {\bf 17} (2000) 3009 [hep-th/9907042];
2572: 
2573: \bibitem{18} 
2574: L. Bombelli, ``Statistical Lorentzian Geometry and the Closeness of 
2575: Lorentzian Manifolds'', J. Math. Phys. {\bf 41} (2000) 
2576: 6944-6958; [gr-qc/0002053]\\
2577: L. Bombelli, A. Corichi, O. Winkler, ``Semiclassical Quantum Gravity:
2578: Statistics of Combinatorial Riemannian Geometries'', [gr-qc/0409006]
2579: 
2580: \bibitem{19} T. Thiemann, O. Winkler, ```Gauge Field Theory Coherent 
2581: States (GCS): 
2582: IV. Infinite Tensor Product and Thermodynamic Limit",
2583: Class. Quantum Grav. {\bf 18} (2001) 4997-5033, [hep-th/0005235]
2584: 
2585: 
2586: \bibitem{10d} T. Thiemann, ``Kinematical Hilbert Spaces for Fermionic and
2587: Higgs Quantum Field Theories",
2588: Class. Quantum Grav. {\bf 15} (1998) 1487-1512, [gr-qc/9705021]\\
2589: A. Ashtekar, J. Lewandowski, H. Sahlmann, ``Polymer and
2590: Fock Representations for a Scalar Field'', Class. Quantum Grav. {\bf 20}
2591: (2003) L11, [gr-qc/0211012]
2592: 
2593: \bibitem{20} 
2594: T. Thiemann, ``Quantum Spin Dynamics (QSD): VII.
2595: Symplectic Structures and Continuum Lattice Formulations of
2596: Gauge Field Theories", Class.Quant.Grav.18:3293-3338,2001,
2597: [hep-th/0005232]; ``Gauge Field Theory Coherent States (GCS): I.
2598: General Properties", Class.Quant.Grav.18:2025-2064,2001, 
2599: [hep-th/0005233]\\
2600: T. Thiemann, O. Winkler, ``Gauge Field Theory Coherent States
2601: (GCS): II. Peakedness Properties", Class.Quant.Grav.18:2561-2636,2001,
2602: [hep-th/0005237]; ``III. Ehrenfest Theorems",
2603: Class. Quantum Grav. {\bf 18} (2001) 4629-4681, [hep-th/0005234]\\
2604: H. Sahlmann, T. Thiemann, O. Winkler, ``Coherent States for
2605: Canonical Quantum General Relativity and the Infinite Tensor Product
2606: Extension", Nucl.Phys.B606:401-440,2001;
2607: [gr-qc/0102038]
2608: 
2609: \bibitem{20b} K. Giesel, T. Thiemann, ``Consistency Check on Volume and 
2610: Triad Operator Quantisation in Loop Quantum Gravity'', to appear
2611: 
2612: \bibitem{21} J. Klauder, ``The Current State of Coherent States'',
2613: quant-ph/0110108 
2614: 
2615: \bibitem{22} 
2616: B. Bolen, L. Bombelli, A. Corichi, ``Semiclassical States in Quantum 
2617: Cosmology: Bianchi One Coherent States''
2618: Class. Quant. Grav. {\bf 21} (2004) 4087-4106; [gr-qc/0404004]
2619: 
2620: \bibitem{22a} T. Thiemann, ``Complexifier Coherent States for Canonical 
2621: Quantum General Relativity", gr-qc/0206037
2622: 
2623: \bibitem{22b} 
2624: K. Noui, A. Perez, K. Vandersloot, ``On the physical Hilbert Space of 
2625: Loop Quantum Cosmology'', Phys. Rev. {\bf D71} (2005) 044025;
2626: [gr-qc/0411039]
2627: 
2628: \bibitem{23} 
2629: S. Tsujikawa, P. Singh, R. Maartens, ``Loop Quantum Gravity Effects on
2630: Inflation and the CMB'', 
2631: Class. Quant. Grav. {\bf 21} (2004) 5767-5775; [astro-ph/0311015]\\
2632: M. Bojowald, R. Maartens, P. Singh, ``Loop Quantum Gravity and the Cyclic
2633: Universe'',  Phys. Rev. {\bf D70} (2004) 083517; [hep-th/0407115]\\
2634: ``Inflation from Quantum Geometry",
2635: Phys. Rev. Lett. {\bf 89} (2002) 261301, [gr-qc/0206054]\\
2636: M. Bojowald, G. Date, ``Quantum Suppression of the Generic Chaotic 
2637: Behaviour to Cosmological Singularities'', Phys. Rev. Lett. 
2638: {\bf 92} (2004) 071302; [gr-qc/0311003]\\
2639: M. Bojowald, ``Initial Conditions for a Universe'', 
2640: Essay awarded first prize in the Gravity Research Foundation 
2641: Essay contest 2003,  Gen. Rel. Grav. {\bf 35} (2003) 1877-1883;
2642: [gr-qc/0305069]
2643: 
2644: 
2645: \end{thebibliography}
2646: 
2647: \end{document}
2648: