gr-qc0505057/elqc.tex
1: \documentclass[12pt]{article}
2: \usepackage{a4wide}
3: \usepackage{amssymb}
4: \usepackage{graphicx}
5: %\voffset-1cm
6: \begin{document}
7: {\renewcommand{\thefootnote}{\fnsymbol{footnote}}
8: \hfill  AEI--2005--025\\ 
9: \medskip
10: \hfill gr--qc/0505057\\
11: \medskip
12: \begin{center}
13: {\LARGE  Elements of Loop Quantum Cosmology}\\
14: \vspace{1.5em}
15: Martin Bojowald\footnote{e-mail address: {\tt mabo@aei.mpg.de}}
16: \\
17: \vspace{0.5em}
18: Max-Planck-Institut f\"ur Gravitationsphysik, Albert-Einstein-Institut,\\
19: Am M\"uhlenberg 1, D-14476 Potsdam, Germany
20: \vspace{1.5em}
21: \end{center}
22: }
23: 
24: \setcounter{footnote}{0}
25: 
26: \newcommand{\lP}{\ell_{\rm P}}
27: \newcommand{\md}{{\rm d}}
28: \newcommand{\sgn}{{\rm sgn}}
29: 
30: \begin{abstract}
31:   The expansion of our universe, when followed backward in time,
32:   implies that it emerged from a phase of huge density, the big bang.
33:   These stages are so extreme that classical general relativity
34:   combined with matter theories is not able to describe them properly,
35:   and one has to refer to quantum gravity. A complete quantization of
36:   gravity has not yet been developed, but there are many results about
37:   key properties to be expected. When applied to cosmology, a
38:   consistent picture of the early universe arises which is free of the
39:   classical pathologies and has implications for the generation of
40:   structure which are potentially observable in the near future.
41: \end{abstract}
42: 
43: \section{Introduction} 
44: 
45: General relativity provides us with an extremely successful
46: description of the structure of our universe on large scales, with
47: many confirmations by macroscopic experiments and so far no conflict
48: with observations.  The resulting picture, when applied to early
49: stages of cosmology, suggests that the universe had a beginning a
50: finite time ago, at a point where space, matter, and also time itself
51: were created. Thus, it does not even make sense to ask what was there
52: before since ``before'' does not exist at all. At very early stages,
53: space was small such that there were huge energy densities to be
54: diluted in the later expansion of the universe that is still
55: experienced today. In order to explain also the structure that we see
56: in the form of galaxies in the correct statistical distribution, the
57: universe not only needs to expand but do so in an accelerated manner,
58: a so-called inflationary period, in its early stages. With this
59: additional input, usually by introducing inflation with exponential
60: acceleration \cite{Guth,NewInfl,InflAS} lasting long enough to expand
61: the scale factor $a(t)$, the radius of the universe at a given time
62: $t$, by a ratio $a_{\rm final}/a_{\rm initial}>e^{60}$. The resulting
63: seeds for structure after the inflationary phase can be observed in
64: the anisotropy spectrum of the cosmic microwave background (most
65: recently of the WMAP satellite \cite{WMAPParam}), which agrees well
66: with theoretical predictions over a large range of scales.
67: 
68: Nonetheless, there are problems remaining with the overall picture.
69: The beginning was extremely violent with conditions such as diverging
70: energy densities and tidal forces under which no theory can prevail.
71: This is also true for general relativity itself which led to this
72: conclusion in the first place: according to the singularity theorems
73: any solution to general relativity, under reasonable conditions on the
74: form of matter, must have a singularity in the past or
75: future \cite{SingTheo}. There, space degenerates, e.g.\ to a single
76: point in cosmology, and energy densities and tidal forces diverge.
77: From the observed expansion of our current universe one can conclude
78: that according to general relativity there must have been such a
79: singularity in the past (which does not rule out further possible
80: singularities in the future). This is exactly what is usually referred
81: to as the ``beginning'' of the universe, but from the discussion it is
82: clear that the singularity does not so much present a boundary to the
83: universe as a boundary to the classical theory: The theory predicts
84: conditions under which it has to break down and is thus incomplete.
85: Here it is important that the singularity in fact lies only a finite
86: time in the past rather than an infinite distance away, which could be
87: acceptable. A definitive conclusion about a possible beginning can
88: therefore be reached only if a more complete theory is found which is
89: able to describe these very early stages meaningfully.
90: 
91: Physically, one can understand the inevitable presence of
92: singularities in general relativity by the characteristic property of
93: classical gravitation being always attractive. In the backward
94: evolution in which the universe contracts, there is, once matter has
95: collapsed to a certain size, simply no repulsive force strong enough
96: to prevent the total collapse into a singularity. A similar behavior
97: happens when not all the matter in the universe but only that in a
98: given region collapses to a small size, leading to the formation of
99: black holes which also are singular.
100: 
101: This is the main problem which has to be resolved before one can call
102: our picture of the universe complete. Moreover, there are other
103: problematic issues in what we described so far. Inflation has to be
104: introduced into the picture, which currently is done by assuming a
105: special field, the inflaton, in addition to the matter we know. In
106: contrast to other matter, its properties must be very exotic so as to
107: ensure accelerated expansion which with Einstein's equations is
108: possible only if there is negative pressure. This is achieved by
109: choosing a special potential and initial conditions for the inflaton,
110: but there is no fundamental explanation of the nature of the inflaton
111: and its properties. Finally, there are some details in the anisotropy
112: spectrum which are hard to bring in agreement with theoretical models.
113: In particular, there seems to be less structure on large scales than
114: expected, referred to as a loss of power.
115: 
116: \section{Classical Cosmology}
117: 
118: In classical cosmology one usually assumes space to be homogeneous and
119: isotropic, which is an excellent approximation on large scales today.
120: The metric of space is then solely determined by the scale factor
121: $a(t)$ which gives the size of the universe at any given time $t$. The
122: function $a(t)$ describes the expansion or contraction of space in a
123: way dictated by the Friedmann equation \cite{Friedmann}
124: \begin{equation} \label{Friedmann}
125:   \left(\frac{\dot{a}}{a}\right)^2=\frac{8\pi}{3}G\rho(a)
126: \end{equation}
127: which is the reduction of Einstein's equations under the assumption of
128: isotropy. In this equation, $G$ is the gravitational constant and
129: $\rho(a)$ the energy density of whatever matter we have in the
130: universe. Once the matter content is chosen and $\rho(a)$ is known,
131: one can solve the Friedmann equation in order to obtain $a(t)$.
132: 
133: As an example we consider the case of radiation which can be
134: described phenomenologically by the energy density $\rho(a)\propto
135: a^{-4}$. This is only a phenomenological description since it ignores
136: the fundamental formulation of electrodynamics of the Maxwell field.
137: Instead of using the Maxwell Hamiltonian in order to define the energy
138: density, which would complicate the situation by introducing the
139: electromagnetic fields with new field equations coupled to the
140: Friedmann equation, one uses the fact that on large scales the energy
141: density of radiation is diluted by the expansion and in addition
142: red-shifted. This leads to a behavior proportional $a^{-3}$ from
143: dilution times $a^{-1}$ from redshift. In this example we then
144: solve the Friedmann equation $\dot{a}\propto a^{-1}$
145: by $a(t)\propto\sqrt{t-t_0}$ with a constant of integration $t_0$.
146: This demonstrates the occurrence of singularities: For any solution
147: there is a time $t=t_0$ where the size of space vanishes and the
148: energy density $\rho(a(t_0))$ diverges. At this point not only the
149: matter system becomes unphysical, but also the gravitational evolution
150: breaks down: When the right hand side of (\ref{Friedmann}) diverges at
151: some time $t_0$, we cannot follow the evolution further by setting up
152: an initial value problem there and integrating the equation. We can
153: thus only learn that there is a singularity in the classical theory,
154: but do not obtain any information as to what is happening there and
155: beyond. These are the two related but not identical features of a
156: singularity: energy densities diverge and the evolution breaks down.
157: 
158: One could think that the problem comes from too strong idealizations
159: such as symmetry assumptions or the phenomenological description of
160: matter. That this is not the case follows from the singularity
161: theorems which do not depend on these assumptions. One can also
162: illustrate the singularity problem with a field theoretic rather than
163: phenomenological description of matter. For simplicity we now assume
164: that matter is provided by a scalar $\phi$ whose energy density then
165: follows from the Hamiltonian
166: \begin{equation} \label{Hscalar}
167:   \rho(a)=a^{-3}H(a)= a^{-3}({\textstyle\frac{1}{2}}a^{-3}p_{\phi}^2+
168: a^3V(\phi))
169: \end{equation}
170: with the scalar momentum $p_{\phi}$ and potential $V(\phi)$. At small
171: scale factors $a$, there still is a diverging factor $a^{-3}$ in the
172: kinetic term which we recognized as being responsible for the
173: singularity before. Since this term dominates over the non-diverging
174: potential term, we still cannot escape the singularity by using this
175: more fundamental description of matter. This is true unless we manage
176: to arrange the evolution of the scalar in such a way that
177: $p_{\phi}\to0$ when $a\to0$ in just the right way for the kinetic term
178: not to diverge. This is difficult to arrange in general, but is
179: exactly what is attempted in slow-roll inflation (though with a
180: different motivation, and not necessarily all the way up to the
181: classical singularity).
182: 
183: For the evolution of $p_{\phi}$ we need the scalar equation of motion,
184: which can be derived from the Hamiltonian $H$ in (\ref{Hscalar}) via
185: $\dot{\phi}=\{\phi,H\}$ and $\dot{p}_{\phi}=\{p_{\phi},H\}$. This
186: results in the isotropic Klein--Gordon equation in a time-dependent
187: background determined by $a(t)$,
188: \begin{equation} \label{KG}
189:  \ddot{\phi}+3\dot{a}a^{-1}\dot{\phi}+V'(\phi)=0\,.
190: \end{equation}
191: In an expanding space with positive $\dot{a}$ the second term implies
192: friction such that, if we assume the potential $V'(\phi)$ to be flat
193: enough, $\phi$ will change only slowly (slow-roll). Thus, $\dot{\phi}$
194: and $p_{\phi}=a^3\dot{\phi}$ are small and at least for some time we
195: can ignore the kinetic term in the energy density. Moreover, since
196: $\phi$ changes only slowly we can regard the potential $V(\phi)$ as a
197: constant $\Lambda$ which again allows us to solve the Friedmann
198: equation with $\rho(a)=\Lambda$. The solution $a\propto
199: \exp(\sqrt{8\pi G\Lambda/3}\,t)$ is inflationary since $\ddot{a}>0$ and
200: non-singular: $a$ becomes zero only in the limit $t\to-\infty$.
201: 
202: Thus, we now have a mechanism to drive a phase of accelerated
203: expansion important for observations of structure. However, this
204: expansion must be long enough, which means that the phase of slowly
205: rolling $\phi$ must be long. This can be achieved only if the
206: potential is very flat and $\phi$ starts sufficiently far away from
207: its potential minimum. Flatness means that the ratio of $V(\phi_{\rm
208:   initial})$ and $\phi_{\rm initial}$ must be of the order $10^{-10}$,
209: while $\phi_{\rm initial}$ must be huge, of the order of the Planck
210: mass \cite{ChaoticInfl}. These assumptions are necessary for agreement
211: with observations, but are in need of more fundamental explanations.
212: 
213: Moreover, inflation alone does not solve the singularity
214: problem \cite{InflSing}. The non-singular solution we just
215: obtained was derived under the approximation that the kinetic term can
216: be ignored when $\dot{\phi}$ is small. This is true in a certain range
217: of $a$, depending on how small $\dot{\phi}$ really is, but never very
218: close to $a=0$. Eventually, even with slow-roll conditions, the
219: diverging $a^{-3}$ will dominate and lead to a singularity.
220: 
221: \section{Quantum Gravity}
222: 
223: For decades, quantum gravity has been expected to complete the
224: picture which is related to well-known properties of quantum mechanics
225: in the presence of a non-zero $\hbar$.
226: 
227: \subsection{Indications}
228: 
229: First, in analogy to the singularity problem in gravity, where
230: everything falls into a singularity in finite time, there is the
231: instability problem of a classical hydrogen atom, where the electron
232: would fall into the nucleus after a brief time. From quantum mechanics
233: we know how the instability problem is solved: There is a finite
234: ground state energy $E_0=-\frac{1}{2} me^4/\hbar^2$, implying that the
235: electron cannot radiate away all its energy and not fall further once
236: it reaches the ground state. From the expression for $E_0$ one can see
237: that quantum theory with its non-zero $\hbar$ is essential for this to
238: happen: When $\hbar\to0$ in a classical limit, $E_0\to-\infty$ which
239: brings us back to the classical instability. One expects a similar
240: role to be played by the Planck length $\lP=\sqrt{8\pi G\hbar/c^3}\approx
241: 10^{-35}{\rm m}$ which is tiny but non-zero in quantum theory. If,
242: just for dimensional reasons, densities are bounded by $\lP^{-3}$,
243: this would be finite in quantum gravity but diverge in the classical
244: limit.
245: 
246: Secondly, a classical treatment of black body radiation suggests the
247: Rayleigh--Jeans law according to which the spectral density behaves as
248: $\rho(\lambda)\propto\lambda^{-4}$ as a function of the wave
249: length. This is unacceptable since the divergence at small wave
250: lengths leads to an infinite total energy. Here, quantum mechanics
251: solves the problem by cutting off the divergence with Planck's formula
252: which has a maximum at a wave length $\lambda_{\rm max}\sim h/kT$ and
253: approaches zero at smaller scales. Again, in the classical limit
254: $\lambda_{\max}$ becomes zero and the expression diverges.
255: 
256: In cosmology the situation is similar for matter in the whole universe
257: rather than a cavity. Energy densities as a function of the scale
258: factor behave as, e.g., $a^{-3}$ if matter is just diluted or $a^{-4}$
259: if there is an additional redshift factor. In all cases, the energy
260: density diverges at small scales, comparable to the Rayleigh--Jeans
261: law. Inflation already provides an indication that the behavior must
262: be different at small scales. Indeed, inflation can only be achieved
263: with negative pressure, while all matter whose energy falls off as
264: $a^{-k}$ with non-negative $k$ has positive pressure. This can easily
265: be seen from the thermodynamical definition of pressure as the
266: negative change of energy with volume. Negative pressure then requires
267: the energy to increase with the scale factor at least at small scales
268: where inflation is required (e.g., an energy $\Lambda a^3$ for
269: exponential inflation). This could be reconciled with standard forms
270: of matter if there is an analog to Planck's formula, which
271: interpolates between decreasing behavior at large scales and a
272: behavior increasing from zero at small scales, with a maximum in
273: between.
274: 
275: \subsection{Early Quantum Cosmology}
276: 
277: Since the isotropic reduction of general relativity leads to a system
278: with finitely many degrees of freedom, one can in a first attempt try
279: quantum mechanics to quantize it. Starting with the Friedmann equation
280: (\ref{Friedmann}) and replacing $\dot{a}$ by its momentum
281: $p_a=3a\dot{a}/8\pi G$ gives a Hamiltonian which is quadratic in
282: the momentum and can be quantized easily to an operator acting on a
283: wave function depending on the gravitational variable $a$ and possibly
284: matter fields $\phi$. The usual Schr\"odinger representation yields
285: the Wheeler--DeWitt equation \cite{DeWitt,QCReview}
286: \begin{equation}\label{WdW}
287:  \frac{3}{2}\left(-\frac{1}{9}\lP^4
288:   a^{-1}\frac{\partial}{\partial a}a^{-1}
289:   \frac{\partial}{\partial a}\right) a
290: \psi(a,\phi)= 8\pi G\hat{H}_{\phi}(a) \psi(a,\phi)
291: \end{equation}
292: with the matter Hamiltonian $\hat{H}_{\phi}(a)$. This system is
293: different from usual quantum mechanics in that there are factor
294: ordering ambiguities in the kinetic term, and that there is no
295: derivative with respect to coordinate time $t$. The latter fact is a
296: consequence of general covariance: the Hamiltonian is a constraint
297: equation restricting allowed states $\psi(a,\phi)$, rather than a
298: Hamiltonian generating evolution in coordinate time.  Nevertheless,
299: one can interpret equation (\ref{WdW}) as an evolution equation in the
300: scale factor $a$, which is then called internal time.  The left hand
301: side thus becomes a second order time derivative, and it means that
302: the evolution of matter is measured relationally with respect to the
303: expansion or contraction of the universe, rather than absolutely in
304: coordinate time.
305: 
306: Straightforward quantization thus gives us a quantum evolution
307: equation, and we can now check what this implies for the singularity.
308: If we look at the equation for $a=0$, we notice first that the matter
309: Hamiltonian still leads to diverging energy densities. If we quantize
310: (\ref{Hscalar}), we replace $p_{\phi}$ by a derivative, but the
311: singular dependence on $a$ does not change; $a^{-3}$ would simply
312: become a multiplication operator acting on the wave function. Moreover,
313: $a=0$ remains a singular point of the quantum evolution equation in
314: internal time. There is nothing from the theory which tells us what
315: physically happens at the singularity or beyond (baring intuitive
316: pictures which have been developed from this
317: perspective \cite{tunneling,nobound}).
318: 
319: So one has to ask what went wrong with our expectations that
320: quantizing gravity should help. The answer is that quantum theory
321: itself did not necessarily fail, but only our simple implementation.
322: Indeed, what we used was just quantum mechanics, while quantum gravity
323: has many consistency conditions to be fulfilled which makes
324: constructing it so complicated. At the time when this formalism was
325: first applied there was in fact no corresponding full quantum theory
326: of gravity which could have guided developments. In such a simple case
327: as isotropic cosmology, most of these consistency conditions
328: trivialize and one can easily overlook important issues. There are
329: many choices in quantizing an unknown system, and tacitly making one
330: choice can easily lead in a wrong direction.
331: 
332: Fortunately, the situation has changed with the development of
333: strong candidates for quantum gravity. This then allows us to
334: reconsider the singularity and other problems from the point of view
335: of the full theory, making sure that also in a simpler cosmological
336: context only those steps are undertaken that have an analog in the full
337: theory.
338: 
339: \subsection{Loop quantum gravity}
340: 
341: Singularities are physically extreme and require special properties of
342: any theory aimed at tackling them. First, there are always strong
343: fields (classically diverging) which requires a non-perturbative
344: treatment. Moreover, classically we expect space to degenerate at the
345: singularity, for instance a single point in an isotropic model. This
346: means that we cannot take the presence of a classical geometry to
347: measure distances for granted, which is technically expressed as
348: background independence. A non-perturbative and background independent
349: quantization of gravity is available in the form of loop quantum
350: gravity \cite{Rov:Loops,ThomasRev,ALRev}, which by now is understood
351: well enough in order to be applicable in physically interesting
352: situations.
353: 
354: Here, we only mention salient features of the theory which will turn
355: out to be important for cosmology. The first one is the kind of basic
356: variables used, which are the Ashtekar
357: connection \cite{AshVar,AshVarReell} describing the curvature of space
358: and a densitized triad describing the metric by a collection of three
359: orthonormal vectors in each point. These variables are important since
360: they allow a background independent representation of the theory,
361: where the connection $A_a^i$ is integrated to holonomies
362: \begin{equation} \label{Hol}
363:  h_e(A)={\cal P}\exp\int_e A_a^i\tau_i \dot{e}^a\md t
364: \end{equation}
365: along curves $e$ in space and the densitized triad $E^a_i$ to fluxes
366: \begin{equation} \label{Flux}
367:  F_S(E)=\int_S E^a_i\tau^in_a\md^2y
368: \end{equation}
369: along surfaces $S$. (In these expressions, $\dot{e}^a$ denotes the
370: tangent vector to a curve and $n_a$ the conormal to a surface, both of
371: which are defined without reference to a background metric. Moreover,
372: $\tau_j=-\frac{1}{2}i\sigma_j$ in terms of Pauli matrices). While
373: usual quantum field theory techniques rest on the presence of a
374: background metric, for instance in order to decompose a field in its
375: Fourier modes and define a vacuum state and particles, this is no
376: longer available in quantum gravity where the metric itself must be
377: turned into an operator. On the other hand, some integration is
378: necessary since the fields themselves are distributional in quantum
379: field theory and do not allow a well-defined representation. This
380: ``smearing'' with respect to a background metric has to be replaced by
381: some other integration sufficient for resulting in honest
382: operators \cite{LoopRep,ALMMT}. This is achieved by the integrations in
383: (\ref{Hol}) and (\ref{Flux}), which similarly lead to a well-defined
384: quantum representation. Usual Fock spaces in perturbative quantum
385: field theory are thereby replaced by the loop representation, where an
386: orthonormal basis is given by spin network states \cite{RS:Spinnet}.
387: 
388: This shows that choosing basic variables for a theory to quantize has
389: implications for the resulting representation. Connections and
390: densitized triads can naturally be smeared along curves and surfaces
391: without using a background metric and then represented on a Hilbert
392: space. Requiring diffeomorphism invariance, which means that a
393: background independent theory must not change under deformations of
394: space (which can be interpreted as changes of coordinates), even
395: selects a unique
396: representation \cite{FluxAlg,Meas,HolFluxRep,SuperSel,WeylRep}. These are
397: basic properties of loop quantum gravity, recognized as important
398: requirements for a background independent quantization. Already here
399: we can see differences to the Wheeler--DeWitt quantization, where the
400: metric is used as a basic variable and then quantized as in quantum
401: mechanics. This is possible in the model but not in a full theory, and
402: in fact we will see later that a loop quantization will give a
403: representation inequivalent to the Wheeler--DeWitt quantization.
404: 
405: The basic properties of the representation have further consequences.
406: Holonomies and fluxes act as well-defined operators, and fluxes have
407: discrete spectra. Since spatial geometry is determined by the
408: densitized triad, spatial geometry is discrete, too, with discrete
409: spectra for, e.g., the area and volume
410: operator \cite{AreaVol,Area,Vol2}. The geometry of space-time is more
411: complicated to understand since this is a dynamical situation which
412: requires solving the Hamiltonian constraint. This is the analog of the
413: Wheeler--DeWitt equation in the full theory and is the quantization of
414: Einstein's dynamical equations. There are candidates for such
415: operators \cite{QSDI}, well-defined even in the presence of
416: matter \cite{QSDV} which in usual quantum field theory would contribute
417: divergent matter Hamiltonians. Not surprisingly, the full situation is
418: hard to analyze, which is already the case classically, without assuming
419: simplifications from symmetries. We will thus return to symmetric, in
420: particular isotropic models, but with the new perspective provided by
421: the full theory of loop quantum gravity.
422: 
423: \section{Quantum cosmology}
424: 
425: Symmetries can be introduced in loop quantum gravity at the level of
426: states and basic operators \cite{PhD,SymmRed,SphSymm}, such that it is not
427: necessary to reduce the classical theory first and then quantize as in
428: the Wheeler--DeWitt quantization. Instead, one can view the procedure
429: as quantizing first and then introducing symmetries which ensures that
430: consistency conditions of quantum gravity are observed in the first
431: step before one considers treatable situations. In particular, the
432: quantum representation derives from symmetric states and basic
433: operators, while the Hamiltonian constraint can be obtained with
434: constructions analogous to those in the full theory. This allows us to
435: reconsider the singularity problem, now with methods from full quantum
436: gravity. In fact, symmetric models present a class of systems which
437: can often be treated explicitly while still being representative for
438: general phenomena. For instance, the prime examples of singular
439: situations in gravity, and some of the most widely studied physical
440: applications, are already obtained in isotropic or spherically
441: symmetric systems, which allow access to cosmology and black holes.
442: 
443: \subsection{Representation}
444: 
445: Before discussing the quantum level we reformulate isotropic cosmology
446: in connection and triad variables instead of $a$. The role of the
447: scale factor is now played by the densitized triad component $p$ with
448: $|p|=a^2$ whose canonical momentum is the isotropic connection
449: component $c=-\frac{1}{2}\dot{a}$ with $\{c,p\}=8\pi G/3$. The main
450: difference to metric variables is the fact that $p$, unlike $a$, can
451: take both signs with $\sgn p$ being the orientation of space. This is
452: a consequence of having to use triad variables which not only know
453: about the size of space but also its orientation (depending on whether
454: the set of orthonormal vectors is left or right handed).
455: 
456: States in the full theory are usually written in the connection
457: representation as functions of holonomies. Following the reduction
458: procedure for an isotropic symmetry group leads to orthonormal states which are
459: functions of the isotropic connection component $c$ and given by \cite{Bohr}
460: \begin{equation} \label{basis}
461:  \langle c|\mu\rangle= e^{i\mu c/2} \qquad \mu\in{\mathbb R}\,.
462: \end{equation}
463: On these states the basic variables $p$ and $c$ are represented by
464: \begin{eqnarray}
465: \hat{p}|\mu\rangle &=& {\textstyle\frac{1}{6}}\lP^2\mu|\mu\rangle\\
466: \widehat{e^{i\mu'c/2}}|\mu\rangle &=& |\mu+\mu'\rangle 
467:  \end{eqnarray}
468: with the properties:
469: \begin{enumerate}
470:  \item $[\widehat{e^{i\mu'
471:        c/2}},\hat{p}]=-\frac{1}{6}\lP^2\mu'\widehat{e^{-i\mu' c/2}}=
472:    i\hbar(\{e^{i\mu' c/2},p\})^{\wedge}$,
473:  \item $\hat{p}$ has a discrete spectrum and
474:  \item only exponentials $e^{i\mu'c/2}$ of $c$ are represented, not
475:    $c$ directly.
476: \end{enumerate}
477: These statements deserve further explanation: First, the classical
478: Poisson relations between the basic variables are indeed represented
479: correctly, turning the Poisson brackets into commutators divided by
480: $i\hbar$.  On this representation, the set of eigenvalues of $\hat{p}$
481: is the full real line since $\mu$ can take arbitrary real values.
482: Nevertheless, the spectrum of $\hat{p}$ is discrete in the technical
483: sense that eigenstates of $\hat{p}$ are normalizable. This is indeed
484: the case in this non-separable Hilbert space where (\ref{basis})
485: defines an orthonormal basis. The last property follows since the
486: exponentials are not continuous in the label $\mu'$, for otherwise one
487: could simply take the derivative with respect to $\mu'$ at $\mu'=0$
488: and obtain an operator for $c$. The discontinuity can be seen, e.g., from
489: \[
490:  \langle\mu|\widehat{e^{i\mu'c/2}}|\mu\rangle= \delta_{0,\mu'}
491: \]
492: which is not continuous.
493: 
494: These properties are quite unfamiliar from quantum mechanics, and
495: indeed the representation is inequivalent to the Schr\"odinger
496: representation (the discontinuity of the $c$-exponential evading the
497: Stone--von Neumann theorem which usually implies uniqueness of the
498: representation). In fact, the loop representation is inequivalent to
499: the Wheeler-DeWitt quantization which just assumed a Schr\"odinger
500: like quantization. In view of the fact that the phase space of our
501: system is spanned by $c$ and $p$ with $\{c,p\}\propto 1$ just as in
502: classical mechanics, the question arises how such a difference in the
503: quantum formulation arises.
504: 
505: As a mathematical problem the basic step of quantization occurs as
506: follows: given the classical Poisson algebra of observables $Q$ and
507: $P$ with $\{Q,P\}=1$, how can we define a representation of the
508: observables on a Hilbert space such that the Poisson relations become
509: commutator relations and complex conjugation, meaning that $Q$ and $P$
510: are real, becomes adjointness? The problem is mathematically much
511: better defined if one uses the bounded expressions $e^{isQ}$ and
512: $e^{it\hbar^{-1}P}$ instead of the unbounded $Q$ and $P$, which still
513: allows us to distinguish any two points in the whole phase space. The
514: basic objects $e^{isQ}$ and $e^{it\hbar^{-1}P}$ upon quantization will
515: then not commute but fulfill the commutation relation (Weyl algebra)
516: \begin{equation}\label{QP}
517:  e^{isQ}e^{it\hbar^{-1}P} = e^{ist} e^{it\hbar^{-1}P} e^{isQ}
518: \end{equation}
519: as unitary operators on a Hilbert space.
520: 
521: In the Schr\"odinger representation this is done by using a Hilbert
522: space $L^2({\mathbb R},\md q)$ of square integrable functions
523: $\psi(q)$ with $\int_{\mathbb R}\md q|\psi(q)|^2$ finite. The
524: representation of basic operators is
525: \begin{eqnarray*}
526:  e^{isQ}\psi(q) &=& e^{isq}\psi(q)\\
527:  e^{it\hbar^{-1}P}\psi(q) &=& \psi(q+t)
528: \end{eqnarray*}
529: which indeed are unitary and fulfill the required commutation
530: relation. Moreover, the operator families as functions of $s$ and $t$
531: are continuous and we can take the derivatives in $s=0$ and $t=0$,
532: respectively:
533: \begin{eqnarray*}
534:  -i\left.\frac{\md}{\md s}\right|_{s=0} e^{isQ} &=& q\\
535:  -i\hbar\left.\frac{\md}{\md t}\right|_{t=0} e^{it\hbar^{-1}P} &=&
536:  \hat{p}=-i\hbar\frac{\md}{\md q}\,.
537: \end{eqnarray*}
538: This is the familiar representation of quantum mechanics which,
539: according to the Stone--von Neumann theorem is unique under the
540: condition that $e^{isQ}$ and $e^{it\hbar^{-1}P}$ are indeed continuous
541: in both $s$ and $t$.
542: 
543: The latter condition is commonly taken for granted in quantum
544: mechanics, but in general there is no underlying physical or
545: mathematical reason. It is easy to define representations violating
546: continuity in $s$ or $t$, for instance if we use a Hilbert space
547: $\ell^2({\mathbb R})$ where states are again maps $\psi_q$ from the
548: real line to complex numbers but with norm $\sum_q|\psi_q|^2$ which
549: implies that normalizable $\psi_q$ can be non-zero for at most
550: countably many $q$. We obtain a representation with basic operators
551: \begin{eqnarray*}
552:  e^{isQ}\psi_q &=& e^{isq}\psi_q\\
553:  e^{it\hbar^{-1}P}\psi_q &=& \psi_{q+t}
554: \end{eqnarray*}
555: which is of the same form as before. However, due to the different
556: Hilbert space the second operator $e^{it\hbar^{-1}P}$ is no longer
557: continuous in $t$ which can be checked as in the case of $e^{i\mu
558:   c/2}$. In fact, the representation for $Q$ and $P$ is isomorphic to
559: that of $p$ and $c$ used before, where a general state
560: $|\psi\rangle=\sum_{\mu}\psi_{\mu}|\mu\rangle$ has coefficients
561: $\psi_{\mu}$ in $\ell^2({\mathbb R})$. 
562: 
563: This explains mathematically why different, inequivalent
564: representations are possible, but what are the physical reasons for
565: using different representations in quantum mechanics and quantum
566: cosmology? In quantum mechanics it turns out that the choice of
567: representation is not that important and is mostly being done for
568: reasons of familiarity with the standard choice. Physical differences
569: between inequivalent representations can only occur at very high
570: energies \cite{PolymerParticle} which are not probed by available
571: experiments and do not affect characteristic quantum effects related
572: to the ground state or excited states. Thus, quantum mechanics as we
573: know it can well be formulated in an inequivalent representation, and
574: also in quantum field theory this can be done and even be
575: useful \cite{QEDBohr}.
576: 
577: In quantum cosmology we have a different situation where it is the high
578: energies which are essential. We do not have direct observations of
579: this regime, but from conceptual considerations such as the
580: singularity issue we have learned which problems we have to face. The
581: classical singularity leads to the highest energies one can imagine,
582: and it is here where the question of which representation to choose
583: becomes essential. As shown by the failure of the Wheeler--DeWitt
584: quantization in trying to remove the singularity, the Schr\"odinger
585: representation is inappropriate for quantum cosmology. The
586: representation underlying loop quantum cosmology, on the other hand,
587: implies very different properties which become important at high
588: energies and can shed new light on the singularity problem.
589: 
590: Moreover, by design of the symmetric models as derived from the full
591: theory, we have the same basic properties of a loop representation in
592: cosmological models and the full situation where they were recognized
593: as being important for a background independent quantization: discrete
594: fluxes $\hat{F}_S(E)$ or $\hat{p}$ and a representation only of
595: holonomies $h_e(A)$ or $e^{i\mu c/2}$ but not of connection components
596: $A_a^i$ or $c$. These basic properties have far-reaching consequences
597: as discussed in what follows \cite{LoopCosRev}:
598: \begin{center}
599: \begin{picture}(160,80)(0,0)
600: \put(80,70){\makebox(0,0){discrete triad \hspace{2cm} only holonomies}}
601: \put(30,60){\vector(0,-1){10}} \put(130,60){\vector(0,-1){10}}
602: \put(80,45){\makebox(0,0){finite inverse volume \hspace{2cm} discrete evolution}}
603: \put(40,40){\vector(1,-1){10}} \put(125,40){\vector(-1,-1){10}}
604: \put(80,25){\makebox(0,0){ non-singular}}
605: \put(30,35){\vector(0,-1){20}} \put(130,35){\vector(0,-1){20}}
606: \put(80,5){\makebox(0,0){\hspace{-1cm}non-perturbative modifications \hspace{1cm}
607: higher order terms}}
608: \end{picture}
609: \end{center}
610: By this reliable quantization of representative and physically
611: interesting models with a known relation to full quantum gravity we
612: are finally able to resolve long-standing issues such as the
613: singularity problem.
614: 
615: \subsection{Quantum evolution}
616: 
617: We will first look at the quantum evolution equation which we obtain as
618: the quantized Friedmann equation. This is modeled on the Hamiltonian
619: constraint of the full theory such that we can also draw some
620: conclusions for the viability of the full constraint.
621: 
622: \subsubsection{Difference equation}
623: 
624: The constraint equation will be imposed on states of the form
625: $|\psi\rangle=\sum_{\mu}\psi_{\mu}|\mu\rangle$ with summation over
626: countably many values of $\mu$. Since the states $|\mu\rangle$ are
627: eigenstates of the triad operator, the coefficients $\psi_{\mu}$ which
628: can also depend on matter fields such as a scalar $\phi$ represent the
629: state in the triad representation, analogous to $\psi(a,\phi)$ before.
630: For the constraint operator we again need operators for the conjugate
631: of $p$, related to $\dot{a}$ in the Friedmann equation. Since this is
632: now the exponential of $c$, which on basis states acts by shifting the
633: label, it translates to a finite shift in the labels of coefficients
634: $\psi_{\mu}(\phi)$. Plugging together all ingredients for a
635: quantization of (\ref{Friedmann}) along the lines of the constraint in
636: the full theory leads to the difference equation \cite{IsoCosmo,Bohr}
637: \begin{eqnarray} \label{DiffEq}
638: &&    (V_{\mu+5}-V_{\mu+3})\psi_{\mu+4}(\phi)- 2
639: (V_{\mu+1}-V_{\mu-1})\psi_{\mu}(\phi)\\\nonumber
640: &&+    (V_{\mu-3}-V_{\mu-5})\psi_{\mu-4}(\phi)
641:   = -{\textstyle\frac{4}{3}}\pi
642: G\ell_{\rm P}^2\hat{H}_{\rm matter}(\mu)\psi_{\mu}(\phi)
643: \end{eqnarray}
644: with volume eigenvalues $V_{\mu}=(\ell_{\rm P}^2|\mu|/6)^{3/2}$
645: obtained from the volume operator $\hat{V}=|\hat{p}|^{3/2}$, and the
646: matter Hamiltonian $\hat{H}_{\rm matter}(\mu)$.
647: 
648: We again have a constraint equation which does not generate evolution
649: in coordinate time but can be seen as evolution in internal time.
650: Instead of the continuous variable $a$ we now have the label $\mu$
651: which only jumps in discrete steps. As for the singularity issue,
652: there is a further difference to the Wheeler--DeWitt equation since
653: now the classical singularity is located at $p=0$ which is in the
654: interior rather than at the boundary of the configuration space.
655: Nevertheless, the classical evolution in the variable $p$ breaks down
656: at $p=0$ and there is still a singularity.  In quantum theory,
657: however, the situation is very different: while the Wheeler--DeWitt
658: equation does not solve the singularity problem, the difference
659: equation (\ref{DiffEq}) uniquely evolves a wave function from some
660: initial values at positive $\mu$, say, to negative $\mu$. Thus, the
661: evolution does not break down at the classical singularity and can
662: rather be continued beyond it. Quantum gravity is thus a theory which
663: is more complete than classical general relativity and is free of
664: limitations set by classical singularities.
665: 
666: An intuitive picture of what replaces the classical singularity can be
667: obtained from considering evolution in $\mu$ as before. For negative
668: $\mu$, the volume $V_{\mu}$ decreases with increasing $\mu$ while
669: $V_{\mu}$ increases for positive $\mu$. This leads to the picture of a
670: collapsing universe before it reaches the classical big bang
671: singularity and re-expands. While at large scales the classical
672: description is good \cite{SemiClass}, when the universe is small close
673: to the classical singularity it starts to break down and has to be
674: replaced by discrete quantum geometry. The resulting quantum
675: evolution does not break down, in contrast to the classical space-time
676: picture which dissolves. Using the fact that the sign of $\mu$, which
677: defines the orientation of space, changes during the transition
678: through the classical singularity one can conclude that the universe
679: turns its inside out during the process. This can have consequences
680: for realistic matter Hamiltonians which violate parity symmetry.
681: 
682: \subsubsection{Meaning of the wave function}
683: \label{sec:wavefct}
684: 
685: An important issue in quantum gravity which is still outstanding even
686: in isotropic models is the interpretation of the wave function and its
687: relation to the problem of time. In the usual interpretation of
688: quantum mechanics the wave function determines probabilities for
689: measurements made by an observer outside the quantum system. Quantum
690: gravity and cosmology, however, are thought of as theories for the
691: quantum behavior of a whole universe such that, by definition, there
692: cannot be an observer outside the quantum system. Accordingly, the
693: question of how to interpret the wave function in quantum cosmology is
694: more complicated. One can avoid the separation into a classical and
695: quantum part of the problem in quantum mechanics by the theory of
696: decoherence which can explain how a world perceived as classical
697: emerges from the fundamental quantum description \cite{Decoherence}.
698: The degree of ``classicality'' is related to the number of degrees of
699: freedom which do not contribute significantly to the evolution but
700: interact with the system nonetheless. Averaging over those degrees of
701: freedom, provided there are enough of them, then leads to a classical
702: picture. This demonstrates why macroscopic bodies under usual
703: circumstances are perceived as classical while in the microscopic
704: world, where a small number of degrees of freedom is sufficient to
705: capture crucial properties of a system, quantum mechanics prevails.
706: This idea has been adapted to cosmology, where a large universe comes
707: with many degrees of freedom such as small inhomogeneities which are
708: not of much relevance for the overall evolution. This is different,
709: however, in a small universe where quantum behavior becomes dominant.
710: 
711: Thus, one can avoid the presence of an observer outside the quantum
712: system. The quantum system is described by its wave function, and in
713: some circumstances one can approximate the evolution by a quantum part
714: being looked at by classical observers within the same
715: system. Properties are then encoded in a relational way: the wave
716: function of the whole system contains information about everything
717: including possible observers. Now, the question has shifted from a
718: conceptual one --- how to describe the system if no outside observers
719: can be available --- to a technical one. One needs to understand how
720: information can be extracted from the wave function and used to
721: develop schemes for intuitive pictures or potentially observable
722: effects. This is particularly pressing in the very early universe
723: where everything including what we usually know as space and time are
724: quantum and no familiar background to lean on is left.
725: 
726: One lesson is that evolution should be thought of as relational by
727: determining probabilities for one degree of freedom under the
728: condition that another degree of freedom has a certain value. If the
729: reference degree of freedom (such as the direction of the hand of a
730: clock) plays a distinguished role for determining the evolution of
731: others, it is called internal time: it is not an absolute time outside
732: the quantum system as in quantum mechanics, and not a coordinate time
733: as in general relativity which could be changed by coordinate
734: transformations. Rather, it is one of the physical degrees of freedom
735: whose evolution is determined by the dynamical laws and which shows
736: how other degrees of freedom change by interacting with them. From
737: this picture it is clear that no external observer is necessary to
738: read off the clock or other measurement devices, such that it is
739: ideally suited to cosmology. What is also clear is that now internal
740: time depends on what we choose it to be, and different questions
741: require different choices. For a lab experiment the hand of a clock
742: would be a good internal time and, when the clock is sufficiently
743: isolated from the physical fields used in the experiment and other
744: outside influence, will not be different from an absolute time except
745: that it is mathematically more complicated to describe. The same
746: clock, on the other hand, will not be good for describing the universe
747: when we imagine to approach a classical singularity. It will simply
748: not withstand the extreme physical conditions, dissolve, and its parts
749: will behave in a complicated irregular manner ill-suited for the
750: description of evolution. Instead, one has to use more global objects
751: which depend on what is going on in the whole universe.
752: 
753: Close to a classical singularity, where one expects monotonic
754: expansion or contraction, the spatial volume of the universe is just
755: the right quantity as internal time. A wave function then determines
756: relationally how matter fields or other gravitational degrees of
757: freedom change with respect to the expansion or contraction of the
758: universe. In our case, this is encoded in the wave function
759: $\psi_{\mu}(\phi)$ depending on internal time $\mu$ (which through the
760: volume defines the size of the universe but also spatial orientation)
761: and matter fields $\phi$. By showing that it is subject to a
762: difference equation in $\mu$ which does not stop at the classical
763: singularity $\mu=0$ we have seen that relational probabilities are
764: defined for all internal times without breaking down anywhere. This
765: shows the absence of singularities and allows developing intuitive
766: pictures, but does not make detailed predictions before relational
767: probabilities are indeed computed and shown how to be observable at
768: least in principle. 
769: 
770: Here, we encounter the main issue in the role of the wave function: we
771: have a relational scheme to understand what the wave function should
772: mean but the probability measure to be used, called the physical inner
773: product, is not known so far. We already used a Hilbert space which we
774: needed to define the basic operators and the quantized Hamiltonian
775: constraint, where wave functions $\psi_{\mu}$, which by definition are
776: non-zero for at most countably many values $\mu\in{\mathbb R}$, have
777: the inner product $\langle\psi|\psi'\rangle=
778: \sum_{\mu}\bar{\psi}_{\mu}\psi'_{\mu}$. This is called the kinematical
779: inner product which is used for setting up the quantum theory. But
780: unlike in quantum mechanics where the kinematical inner product is
781: also used as physical inner product for the probability interpretation
782: of the wave function, in quantum gravity the physical inner product
783: must be expected to be different. This occurs because the quantum
784: evolution equation (\ref{DiffEq}) in internal time is a constraint
785: equation rather than an evolution equation in an external absolute
786: time parameter. Solutions to this constraint in general are not
787: normalizable in the kinematical inner product such that a new physical
788: inner product on the solution space has to be found. There are
789: detailed schemes for a derivation, but despite some
790: progress \cite{Golam,IsoSpinFoam} they are difficult to apply even in
791: isotropic cosmological models and research is still ongoing. An
792: alternative route to extract physical statements will be discussed in
793: Sec.~\ref{sec:pheno} together with the main results.
794: 
795: A related issue, which is also of relevance for the classical limit of
796: the theory is that of oscillations on small scales of the wave
797: function. Being subject to a difference equation means that
798: $\psi_{\mu}$ is not necessarily smooth but can change rapidly when
799: $\mu$ changes by a small amount even when the volume is large. In such
800: a regime one expects classical behavior, but small scale oscillations
801: imply that the wave function is sensitive to the Planck scale.  There
802: are also other issues related to the fact that now a difference rather
803: than differential equation provides the fundamental
804: law \cite{FundamentalDisc}. Before the physical inner product is known
805: one cannot say if these oscillations would imply any effect observable
806: today, but one can still study the mathematical problem of if and when
807: solutions with suppressed oscillations exist. This is easy to answer
808: in the affirmative for isotropic models subject to (\ref{DiffEq})
809: where in some cases one even obtains a unique wave
810: function \cite{DynIn,Essay}. However, already in other homogeneous but
811: anisotropic models the issue is much more complicated to
812: analyze \cite{GenFunc,GenFuncBI}.
813: 
814: In a more general situation than homogeneous cosmology there is an
815: additional complication even if the physical inner product would be
816: known. In general, it is very difficult to find an internal time to
817: capture the evolution of a complicated quantum system, which is called
818: the problem of time in general relativity. In cosmology, the volume is
819: a good internal time to understand the singularity, but it would not
820: be good for the whole history if the universe experiences a recollapse
821: where the volume would not be monotonic. This is even more complicated
822: in inhomogeneous situations such as the collapse of matter into a
823: black hole. Since we used internal time $\mu$ to show how quantum
824: geometry evolves through the classical singularity, it seems that the
825: singularity problem in general cannot be solved before the problem of
826: time is understood. Fortunately, while the availability of an internal
827: time simplifies the analysis, requirements on a good choice can be
828: relaxed for the singularity problem. An internal time provides us with
829: an interpretation of the constraint equation as an evolution equation,
830: but the singularity problem can be phrased independently of this as
831: the problem to extend the wave function on the space of metrics or
832: triads. This implies weaker requirements and also situations can be
833: analyzed where no internal time is known. The task then is to find
834: conditions which characterize a classical singularity, analogous to
835: $p=0$ in isotropic cosmology, and find an evolution parameter which at
836: least in individual parts of an inhomogeneous singularity allows to
837: see how the system can move through it. Inhomogeneous cases are now
838: under study but only partially understood so far, such that in the
839: next section we return to isotropic cosmology.
840: 
841: \subsection{Densities}
842: 
843: In the previous discussion we have not yet mentioned the matter
844: Hamiltonian on the right hand side, which diverges classically and in
845: the Wheeler--DeWitt quantization when we reach the singularity. If
846: this were the case here, the discrete quantum evolution would break
847: down, too. However, as we will see now the matter Hamiltonian does not
848: diverge, which is again a consequence of the loop representation.
849: 
850: \subsubsection{Quantization}
851: 
852: For the matter Hamiltonian we need to quantize the matter field and in
853: quantum gravity also coefficients such as $a^{-3}$ in the kinetic term
854: which now become operators, too. In the Wheeler--DeWitt quantization
855: where $a$ is a multiplication operator, $a^{-3}$ is unbounded and
856: diverges at the classical singularity. In loop quantum cosmology we
857: have the basic operator $\hat{p}$ which one can use to construct a
858: quantization of $a^{-3}$. However, a straightforward quantization
859: fails since, as one of the basic properties, $\hat{p}$ has a discrete
860: spectrum containing zero. In this case, there is no densely defined
861: inverse operator which one could use. This seems to indicate that the
862: situation is even worse: an operator for the kinetic term would not
863: only be unbounded but not even be well-defined. The situation is much
864: better, however, when one tries other quantizations which are more
865: indirect. For non-basic operators such as $a^{-3}$ there are usually
866: many ways to quantize, all starting from the same classical
867: expression. What we can do here, suggested by constructions in the
868: full theory \cite{QSDV}, is to rewrite $a^{-3}$ in a classically
869: equivalent way as
870: \[
871:  a^{-3}=(\pi^{-1} G^{-1}{\rm
872:    tr}\tau_3e^{c\tau_3}\{e^{-c\tau_3},\sqrt{V}\})^6 
873: \]
874: where we only need a readily available positive power of $\hat{p}$.
875: Moreover, exponentials of $c$ are basic operators, where we just used
876: su(2) notation $e^{c\tau_3}=\cos\frac{1}{2}c+2\tau_3\sin\frac{1}{2}c$
877: in order to bring the expression closer to what one would have in the
878: full theory, and the Poisson bracket will become a commutator in
879: quantum theory.
880: 
881: This procedure, after taking the trace, leads to a densely defined
882: operator for $a^{-3}$ despite the nonexistence of an inverse of
883: $\hat{p}$ \cite{InvScale}:
884: \begin{equation}
885:  \widehat{a^{-3}} = \left(8i\lP^{-2} (\sin{\textstyle\frac{1}{2}}c
886:    \sqrt{\hat{V}}
887:  \cos{\textstyle\frac{1}{2}}c - \cos{\textstyle\frac{1}{2}}c
888:  \sqrt{\hat{V}} \sin{\textstyle\frac{1}{2}}c)\right)^6\,.
889: \end{equation}
890: That this operator is indeed finite can be seen from its action on
891: states $|\mu\rangle$ which follows from that of the basic operators:
892: \begin{equation}
893:  \widehat{a^{-3}}|\mu\rangle = \left(4\lP^{-2}
894:  (\sqrt{V_{\mu+1}}-\sqrt{V_{\mu-1}}\,)\right)^6 |\mu\rangle
895: \end{equation}
896: immediately showing the eigenvalues which are all finite. In
897: particular, at $\mu=0$ where we would have the classical singularity
898: the density operator does not diverge but is zero.
899: 
900: This finiteness of densities finally confirms the non-singular
901: evolution since the matter Hamiltonian
902: \begin{equation} \label{HmatterQuant}
903:  \hat{H}_{\rm
904:    matter}={\textstyle\frac{1}{2}}\widehat{a^{-3}}\hat{p}_{\phi}^2+ 
905:   \hat{V} V(\phi)
906: \end{equation}
907: in the example of a scalar is well-defined even on the classically
908: singular state $|0\rangle$. The same argument applies for other
909: matter Hamiltonians since only the general structure of kinetic and
910: potential terms is used.
911: 
912: \subsubsection{Confirmation of indications}
913: 
914: The finiteness of the operator is a consequence of the loop
915: representation which forced us to take a detour in quantizing inverse
916: powers of the scale factor. A more physical understanding can be
917: obtained by exploiting the fact that there are quantization
918: ambiguities in this non-basic operator. This comes from the rewriting
919: procedure which is possible in many classically equivalent ways,
920: which all lead to different operators. Important properties such as
921: the finiteness and the approach to the classical limit at large volume
922: are robust under the ambiguities, but finer details can change. The
923: most important choices one can make are selecting the representation
924: $j$ of SU(2) holonomies before taking the trace \cite{Gaul,Ambig} and
925: the power $l$ of $|p|$ in the Poisson bracket \cite{ICGC}. These
926: values are restricted by the requirement that $j$ is a half-integer
927: ($j=1/2$ in the above choice) and $0<l<1$ to obtain a well-defined
928: inverse power of $a$ ($l=3/4$ above). The resulting eigenvalues can be
929: computed explicitly and be approximated by the formula \cite{Ambig,ICGC}
930: \begin{equation} \label{effdens}
931:  (a^{-3})_{\rm eff}= a^{-3} p_l(a^2/a_{\rm
932:   max}^2)^{3/(2-2l)}
933: \end{equation}
934: where $a_{\rm max}=\sqrt{j/3}\,\ell_{\rm P}$ depends on the first
935: ambiguity parameter and the function
936: \begin{eqnarray}
937:  p_l(q) &=&
938: \frac{3}{2l}q^{1-l}\left((l+2)^{-1}
939: \left((q+1)^{l+2}-|q-1|^{l+2}\right)\right. \label{plq}\\
940:  &&\qquad- \left.(l+1)^{-1}q
941: \left((q+1)^{l+1}-{\rm sgn}(q-1)|q-1|^{l+1}\right)\right) \,. \nonumber
942: \end{eqnarray}
943: on the second.
944: 
945: \begin{figure}[th]              
946: \centerline{\includegraphics[width=14cm]{pl.eps}}
947: \vspace*{8pt}
948: \caption{The function $p_l(q)$ in (\ref{plq}) for some values of $l$,
949:   including the limiting cases $l=0$ and $l=1$.
950: \label{pl}}
951: \end{figure}
952: 
953: The function $p_l(q)$, shown in Fig.~\ref{pl}, approaches one for
954: $q\gg1$, has a maximum close to $q=1$ and drops off as $q^{2-l}$ for
955: $q\ll1$. This shows that $(a^{-3})_{\rm eff}$ approaches the classical
956: behavior $a^{-3}$ at large scales $a\gg a_{\rm max}$, has a maximum
957: around $a_{\rm max}$ and falls off like $(a^{-3})_{\rm eff}\sim
958: a^{3/(1-l)}$ for $a\ll a_{\rm max}$. The peak value can be
959: approximated, e.g.\ for $j=1/2$, by $(a^{-3})_{\rm eff}(a_{\rm
960:   max})\sim 3l^{-1}2^{-l}(1-3^{-l})^{3/(2-2l)} \lP^{-3}$ which indeed
961: shows that densities are bounded by inverse powers of the Planck
962: length such that they are finite in quantum gravity but diverge in the
963: classical limit.  This confirms our qualitative expectations from the
964: hydrogen atom, while details of the coefficients depend on the
965: quantization.
966: 
967: Similarly, densities are seen to have a peak at $a_{\rm max}$ whose
968: position is given by the Planck length (and an ambiguity
969: parameter). Above the peak we have the classical behavior of an
970: inverse power, while below the peak the density increases from
971: zero. As suggested by the behavior of radiation in a cavity whose
972: spectral energy density
973: \[
974:  \rho_T(\lambda)=8\pi h\lambda^{-5}(e^{h/kT\lambda}-1)^{-1} = h
975:  \lambda^{-5} f(\lambda/\lambda_{\rm max})
976: \]
977: can, analogously to (\ref{effdens}), be expressed as the diverging
978: behavior $\lambda^{-5}$ multiplied with a cut-off function
979: $f(y)=8\pi/((5/(5-x))^{1/y}-1)$ with $x=5+W(-5e^{-5})$ (in terms of
980: the Lambert function $W(x)$, the inverse function of $xe^x$) and
981: $\lambda_{\rm max}=h/xkT$, we obtain an interpolation between
982: increasing behavior at small scales and decreasing behavior at large
983: scales in such a way that the classical divergence is cut off.
984: 
985: \begin{figure}[th]              
986: \centerline{\includegraphics[width=15cm]{Density.eps}}
987: %\vspace*{8pt}
988: \caption{Comparison between the spectral energy density of black body
989:   radiation (wide curve) and an effective geometrical density with
990:   their large scale approximations (dashed).
991: \label{Density}}
992: \end{figure}
993: 
994: We thus have an interpolation between increasing behavior necessary
995: for negative pressure and inflation and the classical decreasing
996: behavior (Fig.~\ref{Density}). Any matter density turns to increasing
997: behavior at sufficiently small scales without the need to introduce an
998: inflaton field with tailor-made properties. In the following section
999: we will see the implications for cosmological evolution by studying
1000: effective classical equations incorporating this characteristic loop
1001: effect of modified densities at small scales.
1002: 
1003: \subsection{Phenomenology}
1004: \label{sec:pheno}
1005: 
1006: The quantum difference equation (\ref{DiffEq}) is rather complicated
1007: to study in particular in the presence of matter fields and, as
1008: discussed in Sec.~\ref{sec:wavefct}, difficult to interpret in a fully
1009: quantum regime. It is thus helpful to work with effective equations,
1010: comparable conceptually to effective actions in field theories, which
1011: are easier to handle and more familiar to interpret but still show
1012: important quantum effects. This can be done
1013: systematically \cite{Bohr,Perturb,Josh}, starting with the Hamiltonian
1014: constraint operator, resulting in different types of correction terms
1015: whose significance in given regimes can be estimated or studied
1016: numerically \cite{Time}. There are perturbative corrections to the
1017: Friedmann equation of higher order form in $\dot{a}$, or of higher
1018: derivative, in the gravitational part on the left hand side, but also
1019: modifications in the matter Hamiltonian since the density in its
1020: kinetic term behaves differently at small scales. The latter
1021: corrections are mainly non-perturbative since the full expression for
1022: $(a^{-3})_{\rm eff}$ depends on the inverse Planck length, and their
1023: range can be extended if the parameter $j$ is rather large. For these
1024: reasons, those corrections are most important and we focus on them
1025: from now on.
1026: 
1027: The effective Friedmann equation then takes the form
1028: \begin{equation} \label{effFried}
1029:   a\dot{a}^2={\textstyle\frac{8\pi}{3}}G
1030: \left({\textstyle\frac{1}{2}}(a^{-3})_{\rm eff}\, p_{\phi}^2+a^3
1031: V(\phi)\right)
1032: \end{equation}
1033: with $(a^{-3})_{\rm eff}$ as in (\ref{effdens}) with a choice of
1034: ambiguity parameters. Since the matter Hamiltonian does not just act
1035: as a source for the gravitational field on the right hand side of the
1036: Friedmann equation, but also generates Hamiltonian equations of
1037: motion, the modification entails further changes in the matter
1038: equations of motion. The Klein--Gordon equation (\ref{KG}) then takes
1039: the effective form
1040: \begin{equation} \label{effKG}
1041:  \ddot{\phi}=\dot{\phi}\,\dot{a}\frac{\md\log(a^{-3})_{\rm
1042: eff}}{\md a}-a^3(a^{-3})_{\rm eff}V'(\phi)
1043: \end{equation}
1044: and finally there is the Raychaudhuri equation
1045: \begin{equation} \label{effRay}
1046:  \frac{\ddot{a}}{a}= -\frac{8\pi G}{3}\left( a^{-3}d(a)_{\rm
1047: eff}^{-1}\dot{\phi}^2 
1048: \left(1-{\textstyle\frac{1}{4}}a\frac{\md
1049: \log(a^3d(a)_{\rm eff})}{\md a}\right) -V(\phi)\right)
1050: \end{equation}
1051: which follows from the above equation and the continuity equation of
1052: matter.
1053: 
1054: \subsubsection{Bounces}
1055: 
1056: The resulting equations can be studied numerically or with qualitative
1057: analytic techniques. We first note that the right hand side of
1058: (\ref{effFried}) behaves differently at small scales since it
1059: increases with $a$ at fixed $\phi$ and $p_{\phi}$. Viewing this
1060: equation as analogous to a constant energy equation in classical
1061: mechanics with kinetic term $\dot{a}^2$ and potential term ${\cal
1062:   V}(a):=-{\textstyle\frac{8\pi}{3}}G a^{-1}
1063: \left({\textstyle\frac{1}{2}}(a^{-3})_{\rm eff}\, p_{\phi}^2+a^3
1064:   V(\phi)\right)$ illustrates the classically attractive nature of
1065: gravity: The dominant part of this potential behaves like $-a^{-4}$
1066: which is increasing. Treating the scale factor analogously to the
1067: position of a classical particle shows that $a$ will be driven toward
1068: smaller values, implying attraction of matter and energy in the
1069: universe. This changes when we approach smaller scales and take into
1070: account the quantum modification. Below the peak of the effective
1071: density the classical potential ${\cal V}(a)$ will now decrease,
1072: $-{\cal V}(a)$ behaving like a positive power of $a$. This implies
1073: that the scale factor will be repelled away from $a=0$ such that there
1074: is now a small-scale repulsive component to the gravitational force if
1075: we allow for quantum effects. The collapse of matter can then be
1076: prevented if repulsion is taken into account, which indeed can be
1077: observed in some models where the effective classical equations alone
1078: are sufficient to demonstrate singularity-free evolution.
1079: 
1080: This happens by the occurrence of bounces where $a$ turns around from
1081: contracting to expanding behavior. Thus, $\dot{a}=0$ and $\ddot{a}>0$.
1082: The first condition is not always realizable, as follows from the
1083: Friedmann equation (\ref{Friedmann}). In particular, when the scalar
1084: potential is non-negative there is no bounce, which is not changed by
1085: the effective density. There are then two possibilities for bounces in
1086: isotropic models, the first one if space has positive curvature rather
1087: than being flat as assumed here \cite{BounceClosed,BounceQualitative},
1088: the second one with a scalar potential which can become
1089: negative \cite{Oscill,Cyclic}. Both cases allow $\dot{a}=0$
1090: even in the classical case, but this always corresponds to a maximum
1091: rather than minimum. This can easily be seen for the case of negative
1092: potential from the Raychaudhuri equation (\ref{effRay}) which in the
1093: classical case implies negative $\ddot{a}$. With the modification,
1094: however, the additional term in the equation provides a positive
1095: contribution which can become large enough for $\ddot{a}$ to become
1096: positive at a value of $\dot{a}=0$ such that there is a bounce.
1097: 
1098: This provides intuitive explanations for the absence of singularities
1099: from quantum gravity, but not a general one. The generic presence of
1100: bounces depends on details of the model such as its matter content or
1101: which correction terms are being used \cite{EffHam,GenericBounce}, and
1102: even with the effective modifications there are always models which
1103: classically remain singular. Thus, the only general argument for
1104: absence of singularities remains the quantum one based on the
1105: difference equation (\ref{DiffEq}), where the conclusion is model
1106: independent.
1107: 
1108: \subsubsection{Inflation}
1109: 
1110: A repulsive contribution to the gravitational force can not only
1111: explain the absence of singularities, but also enhances the expansion
1112: of the universe on scales close to the classical singularity. Thus, as
1113: seen also in Fig.~\ref{InflAeff} the universe accelerates just from
1114: quantum effects, providing a mechanism for inflation without choosing
1115: special matter.
1116: 
1117: \begin{figure}[th]              
1118: \centerline{\includegraphics[width=16cm]{InflAeff.eps}}
1119: \vspace*{8pt}
1120: \caption{Numerical solution to the effective Friedmann equation
1121:   (\ref{effFried}) with a vanishing scalar potential. While the
1122:   modification in the density on the left is active the expansion is
1123:   accelerated, which stops automatically once the universe expands to
1124:   a size above the peak in the effective density.
1125: \label{InflAeff}}
1126: \end{figure}
1127: 
1128: Via the generation of structure, inflationary phases of the universe
1129: can have an imprint on the observable cosmic microwave background.
1130: Observations imply that the predicted power spectrum of anisotropies
1131: must be nearly independent of the scale on which the anisotropies are
1132: probed, which implies that the inflationary phase responsible for
1133: structure formation must be close to exponential acceleration. This is
1134: true for slow-roll inflation, but also for the inflationary phase
1135: obtained from the effective density once a non-zero scalar potential
1136: is taken into account \cite{GenericInfl}. For more detailed comparisons
1137: between theory and observations one needs to consider how
1138: inhomogeneous fields evolve, which already requires us to relax the
1139: strong symmetry assumption of homogeneity. The necessary methods are
1140: not well-developed at the current stage (see
1141: \cite{SphSymm,Horizon,SphSymmHam} for the basic
1142: formulation), but preliminary calculations of implications on the
1143: power spectrum have been undertaken nonetheless.
1144: Ref.~\cite{PowerLoop} indicates that loop inflation can be
1145: distinguished from simple inflaton models because the power depends
1146: differently on scales. 
1147: 
1148: It turns out that this loop phase alone can provide a sufficient
1149: amount of inflation only for unnatural choices of parameters (such as
1150: extremely large $j$), and those cases are even ruled out by
1151: observations already. At this point, the modified matter dynamics of
1152: (\ref{effKG}) and its $\dot{\phi}$-term becomes important.
1153: Classically, it is a friction term which is used for slow-roll
1154: inflation. But in the modified regime at small scales the sign of the
1155: term changes once $(a^{-3})_{\rm eff}$ is increasing. Thus, at those
1156: small scales friction turns into antifriction and matter is driven up
1157: its potential if it has a non-zero initial momentum (even a tiny one,
1158: e.g., from quantum fluctuations). After the loop phase matter fields
1159: slow down and roll back toward their minima, driving additional
1160: inflation. The potentials need not be very special since structure
1161: formation in the first phase and providing a large universe happen by
1162: different mechanisms. When matter fields reach their minima they start
1163: to oscillate and usual re-heating to obtain hot matter can commence.
1164: 
1165: Loop quantum cosmology thus presents a viable alternative to usual
1166: inflaton models which is not in conflict with current observations
1167: but can be verified or ruled out with future experiments such as the
1168: Planck satellite. Its attractive feature is that it does not require
1169: the introduction of an inflaton field with its special properties, but
1170: provides a fundamental explanation of acceleration from quantum
1171: gravity. This scenario is thus encouraging, but so far has not been
1172: developed to the same extent as inflaton models.
1173: 
1174: Even if we assume the presence of an inflaton field are its properties
1175: less special than in the purely classical treatment. We still need to
1176: assume a potential which is sufficiently flat, but there is now an
1177: explanation of initial values far away from the minimum. For this we
1178: again use the effective Klein--Gordon equation and the fact that
1179: $\phi$ is driven up its potential. One can then check that for usual
1180: inflaton potentials the value of typical initial conditions, as a
1181: function of chosen ambiguity parameters and initial fluctuations of
1182: the scalar, is just what one needs for sufficient inflation in a wide
1183: range \cite{InflationWMAP,Robust}. After the modifications in the
1184: density subside, the inflaton keeps moving up the potential from its
1185: initial push, but is now slowed down by the friction term. Eventually,
1186: it will stop and turn around, entering a slow roll phase in its
1187: approach to the potential minimum.  Thus, the whole history of the
1188: expansion is described by a consistent model as illustrated in
1189: Fig.~\ref{Push}, not just the slow roll phase after the inflaton has
1190: already obtained its large initial values.
1191: 
1192: \begin{figure}
1193: \centerline{\includegraphics[width=17cm]{Push.eps}}
1194: \caption{History of the scale factor (top) and inflaton (bottom) with
1195:   the left hand side in slow motion. (Tics on the right horizontal
1196:   axis mark increments in $t$ by 100.) The upper right data are
1197:    rescaled so as to fit on the same plot. Units for $a$ and $\phi$
1198:    are Planck units, and parameters are not necessarily realistic but
1199:    chosen for plotting purposes. Dashed lines mark
1200:   the time and scale factor where classical behavior of $(a^{-3})_{\rm
1201:     eff}$ starts.
1202: \label{Push}} 
1203: \end{figure}
1204: 
1205: One may think that such a second phase of slow-roll inflation washes
1206: away potential quantum gravity effects from the early expansion. That
1207: this is not necessarily the case has been shown in
1208: \cite{InflationWMAP}, based on the fact that around the turning point
1209: of the inflaton the slow-roll conditions are violated.  In this
1210: scenario, structure we see today on the largest scales was created at
1211: the earliest stages of the second inflationary phase since it was
1212: enlarged by the full inflationary phase. If the second inflationary
1213: regime did not last too long, these scales are just observable today
1214: where in fact the observed loss of power can be taken as an indication
1215: of early violations of slow-roll expansion. Thus, loop quantum
1216: cosmology can provide an explanation, among others, for the
1217: suppression of power on large scales.
1218: 
1219: There are diverse scenarios since different phases of inflation can be
1220: combined, and eventually the right one has to be determined from
1221: observations. One can also combine bounces and inflationary regimes in
1222: order to obtain cyclic universes which eventually reach a long phase
1223: of accelerated expansion \cite{InflOsc}. In particular, this allows the
1224: conctruction of models which start close to a simple, static initial
1225: state and, after a series of cycles, automatically reach values of the
1226: scalar to start inflation. In this way, a semiclassical non-singular
1227: inflationary model \cite{Emergent,Emergent2,EmergentLoop} is formulated
1228: which evades the singularity theorem of \cite{InflSing}.
1229: 
1230: Current observations are already beginning to rule out
1231: certain, very large values of the ambiguity parameter $j$ such that
1232: from future data one can expect much tighter limits. In all
1233: these scenarios the non-perturbative modification of the density is
1234: important, which is a characteristic feature of loop quantum
1235: cosmology. At larger scales above the peak there are also perturbative
1236: corrections which imply small changes in the cosmological expansion
1237: and the evolution of field modes. This has recently been
1238: investigated \cite{PowerPert} with the conclusion that potential
1239: effects on the power spectrum would be too small to be noticed by the
1240: next generation satellites. The best candidates for observable effects
1241: from quantum gravity thus remain the non-perturbative modifications in
1242: effective densities.
1243: 
1244: \section{Conclusions and Outlook}
1245: 
1246: What we have described is a consistent picture of the universe which
1247: is not only observationally viable but also mathematically
1248: well-defined and non-singular. There are instances where quantum
1249: gravity is essential, and others where it is helpful in achieving
1250: important effects. The background independent quantization employed
1251: here is very efficient: There are a few basic properties, such as the
1252: discreteness of spatial geometry and the representation only of
1253: exponentials of curvature, which are behind a variety of applications.
1254: Throughout all the developments, those properties have been known to
1255: be essential for mathematical consistency before they were recognized
1256: as being responsible for physical phenomena.
1257: 
1258: For instance, for the singularity issue the basic properties are all
1259: needed in the way they turn out to be realized. First, the theory had
1260: to be based on densitized triad variables which now not only provides
1261: us with the sign of orientation, and thus two sides of the classical
1262: singularity, but also in more complicated models positions the
1263: classical singularity in phase space such that it becomes accessible
1264: by quantum evolution.  Then, the discreteness of spatial geometry
1265: encoded in triad operators and the representation of exponentials of
1266: curvature play together in the right way to remove divergences in
1267: densities and extend the quantum evolution through the classical
1268: singularity. These features allow general results about the absence of
1269: singularities without any new or artificial ingredients, and lead to a
1270: natural solution of a long-standing problem which has eluded previous
1271: attempts for decades.  Symmetry assumptions are still important in
1272: order to be able to perform the calculations, but they can now be
1273: weakened considerably and are not responsible for physical
1274: implications. The essential step is to base the symmetry reduction on
1275: a candidate for full quantum gravity which is background independent
1276: so as to allow studying quantum geometry purely.
1277: 
1278: Absence of singularities in this context is a rather general
1279: statement about the possibility to extend a quantum wave function
1280: through a regime which classically would appear as a singularity. More
1281: explicit questions, such as what kind of new region one is evolving to
1282: and whether it again becomes classical or retains traces of the
1283: evolution through a quantum regime, depend on details of the relevant
1284: constraint operators. This includes, for instance, quantization
1285: ambiguities and the question whether a symmetric operator has to
1286: be used. The latter aspect is also important for technical concepts
1287: such as a physical inner product.
1288: 
1289: Here we discussed only isotropic models which are classically
1290: described solely by the scale factor determining the size of space.
1291: But a more realistic situation has to take into account also the shape
1292: of space, and changes of the distribution of geometry and matter
1293: between different points of space. The methods we used have been
1294: extended to homogeneous models, allowing for anisotropic spaces, and
1295: recently to some inhomogeneous ones, defined by spherical symmetry and
1296: some forms of cylindrical symmetry. In all cases, essential aspects of
1297: the general mechanism for removing classical singularities which has
1298: first been seen only in the simple isotropic models are known to be
1299: realized.\footnote{This does not refer to the boundedness of densities
1300:   or curvature components for {\em all geometries}, which is known not
1301:   to be present in anisotropic models \cite{HomCosmo,Spin} or even on
1302:   some degenerate configurations in the full theory \cite{BoundFull}.
1303:   What is relevant is the behavior on configurations seen along the
1304:   dynamical evolution.}  Moreover, in the more complicated systems it
1305: is acting much more non-trivially, again with the right ingredients
1306: provided by the background independent quantization. Nevertheless,
1307: since the inhomogeneous constraints are much more complicated to
1308: analyze, absence of singularities for them has not yet been proven
1309: completely.  The inhomogeneous systems now also allow access to black
1310: hole and gravitational wave models such that their quantum geometry
1311: can be studied, too.
1312: 
1313: Effective equations are a useful tool to study quantum effects in a
1314: more familiar setting given by classical equations of motion. They
1315: show diverse effects whose usefulness in cosmological phenomenology is
1316: often surprising. Also here, the effects were known to occur from the
1317: quantization and the transfer into effective classical equations,
1318: before they turned out to be helpful. In addition to inflationary
1319: scenarios and bounces which one can see in isotropic cosmologies,
1320: modified densities have more implications in less symmetric models.
1321: The anisotropic Bianchi IX model, for instance, is classically chaotic
1322: which is assumed to play a role in the complicated approach to a
1323: classical singularity \cite{BKL}. With the effective modifications the
1324: dynamics changes and simplifies, removing the classical
1325: chaos \cite{NonChaos}. This has implications for the effective approach
1326: to a classical singularity and can provide a more consistent picture
1327: of general singularities \cite{ChaosLQC}. Effective classical equations
1328: can also be used to study the collapse of matter to a black hole, with
1329: modifications in the development of classical singularities and
1330: horizons \cite{Collapse}. This can now also be studied with
1331: inhomogeneous quantum models which allow new applications for black
1332: holes and cosmological phenomenology where the evolution of
1333: inhomogeneities is of interest in the context of structure formation.
1334: 
1335: With these models there will be new effects not just in cosmology but
1336: also for black holes and other systems which further check the overall
1337: consistency of the theory. Moreover, a better understanding of
1338: inhomogeneities evolving in a cosmological background will give us a
1339: much better computational handle on signatures in the cosmic microwave
1340: or even gravitational wave background, which may soon be testable with
1341: a new generation of observations. One may wonder how it can be
1342: possible to observe quantum gravity effects, given that the Planck
1343: scale is so many orders of magnitude away from scales accessible by
1344: today's experiments. The difference in scales, however, does not
1345: preclude the observation of indirect effects even though direct
1346: measurements on the discreteness scale are impossible, as illustrated
1347: by a well-known example: Brownian motion allows to draw conclusions
1348: about the atomic structure of matter and its size by observations on
1349: much larger scales \cite{Brownian}. Similarly, cosmological
1350: observations can carry information on quantum gravity effects which
1351: otherwise would manifest themselves only at the Planck scale.
1352: 
1353: %\bibliographystyle{../preprint}
1354: %\bibliography{../Bib/QuantGra}
1355: 
1356: \begin{thebibliography}{10}
1357: 
1358: \bibitem{Guth}
1359: A.~H.\ Guth,
1360: \newblock The inflationary universe: A possible solution to the horizon and
1361:   flatness problems,
1362: \newblock {\em Phys.\ Rev.\ D} 23 (1981) 347--356
1363: 
1364: \bibitem{NewInfl}
1365: A.~D.\ Linde,
1366: \newblock A new inflationary universe scenario: A possible solution of the
1367:   horizon, flatness, homogeneity, isotropy and primordial monopole problems,
1368: \newblock {\em Phys.\ Lett.\ B} 108 (1982) 389--393
1369: 
1370: \bibitem{InflAS}
1371: A.\ Albrecht and P.~J.\ Steinhardt,
1372: \newblock Cosmology for grand unified theories with radiatively induced
1373:   symmetry breaking,
1374: \newblock {\em Phys.\ Rev.\ Lett.} 48 (1982) 1220--1223
1375: 
1376: \bibitem{WMAPParam}
1377: D.~N.\ Spergel et~al.,
1378: \newblock First year Wilkinson Microwave Anisotropy Probe (WMAP) observations:
1379:   Determination of cosmological parameters,
1380: \newblock {\em Astrophys.\ J.\ Suppl.} 148 (2003) 175, [astro-ph/0302209]
1381: 
1382: \bibitem{SingTheo}
1383: S.~W.\ Hawking and R.\ Penrose,
1384: \newblock The singularities of gravitational collapse and cosmology,
1385: \newblock {\em Proc.\ Roy.\ Soc.\ Lond.\ A} 314 (1970) 529--548
1386: 
1387: \bibitem{Friedmann}
1388: A.~Friedmann,
1389: \newblock \"Uber die Kr\"ummung des Raumes,
1390: \newblock {\em Z.\ Phys.} 10 (1922) 377--386
1391: 
1392: \bibitem{ChaoticInfl}
1393: A.~D.\ Linde,
1394: \newblock Chaotic Inflation,
1395: \newblock {\em Phys.\ Lett.\ B} 129 (1983) 177--181
1396: 
1397: \bibitem{InflSing}
1398: A.\ Borde, A.~H.\ Guth, and A.\ Vilenkin,
1399: \newblock Inflationary spacetimes are not past-complete,
1400: \newblock {\em Phys.\ Rev.\ Lett.} 90 (2003) 151301, [gr-qc/0110012]
1401: 
1402: \bibitem{DeWitt}
1403: B.~S.\ DeWitt,
1404: \newblock Quantum Theory of Gravity. I. The Canonical Theory,
1405: \newblock {\em Phys.\ Rev.} 160 (1967) 1113--1148
1406: 
1407: \bibitem{QCReview}
1408: D.~L.\ Wiltshire,
1409: \newblock An introduction to quantum cosmology,
1410: \newblock In B.\ Robson, N.\ Visvanathan, and W.~S.\ Woolcock, editors, {\em
1411:   Cosmology: The Physics of the Universe}, pages 473--531. World Scientific,
1412:   Singapore, 1996, [gr-qc/0101003]
1413: 
1414: \bibitem{tunneling}
1415: A.\ Vilenkin,
1416: \newblock Quantum creation of universes,
1417: \newblock {\em Phys.\ Rev.\ D} 30 (1984) 509--511
1418: 
1419: \bibitem{nobound}
1420: J.~B.\ Hartle and S.~W.\ Hawking,
1421: \newblock Wave function of the Universe,
1422: \newblock {\em Phys.\ Rev.\ D} 28 (1983) 2960--2975
1423: 
1424: \bibitem{Rov:Loops}
1425: C.\ Rovelli,
1426: \newblock Loop Quantum Gravity,
1427: \newblock {\em Living Reviews in Relativity} 1 (1998) 1, [gr-qc/9710008],
1428: \newblock http://www.livingreviews.org/Articles/Volume1/1998-1rovelli
1429: 
1430: \bibitem{ThomasRev}
1431: T.\ Thiemann,
1432: \newblock Introduction to Modern Canonical Quantum General Relativity,
1433:   [gr-qc/0110034]
1434: 
1435: \bibitem{ALRev}
1436: A.\ Ashtekar and J.\ Lewandowski,
1437: \newblock Background independent quantum gravity: A status report,
1438: \newblock {\em Class.\ Quantum Grav.} 21 (2004) R53--R152, [gr-qc/0404018]
1439: 
1440: \bibitem{AshVar}
1441: A.\ Ashtekar,
1442: \newblock New Hamiltonian Formulation of General Relativity,
1443: \newblock {\em Phys.\ Rev.\ D} 36 (1987) 1587--1602
1444: 
1445: \bibitem{AshVarReell}
1446: J.~F.\ Barbero~G.,
1447: \newblock Real Ashtekar Variables for Lorentzian Signature Space-Times,
1448: \newblock {\em Phys.\ Rev.\ D} 51 (1995) 5507--5510, [gr-qc/9410014]
1449: 
1450: \bibitem{LoopRep}
1451: C.\ Rovelli and L.\ Smolin,
1452: \newblock Loop Space Representation of Quantum General Relativity,
1453: \newblock {\em Nucl.\ Phys.\ B} 331 (1990) 80--152
1454: 
1455: \bibitem{ALMMT}
1456: A.\ Ashtekar, J.\ Lewandowski, D.\ Marolf, J.\ Mour\~ao, and T.\ Thiemann,
1457: \newblock Quantization of Diffeomorphism Invariant Theories of Connections with
1458:   Local Degrees of Freedom,
1459: \newblock {\em J.\ Math.\ Phys.} 36 (1995) 6456--6493, [gr-qc/9504018]
1460: 
1461: \bibitem{RS:Spinnet}
1462: C.\ Rovelli and L.\ Smolin,
1463: \newblock Spin Networks and Quantum Gravity,
1464: \newblock {\em Phys.\ Rev.\ D} 52 (1995) 5743--5759
1465: 
1466: \bibitem{FluxAlg}
1467: H.\ Sahlmann,
1468: \newblock Some Comments on the Representation Theory of the Algebra Underlying
1469:   Loop Quantum Gravity, [gr-qc/0207111]
1470: 
1471: \bibitem{Meas}
1472: H.\ Sahlmann,
1473: \newblock When Do Measures on the Space of Connections Support the Triad
1474:   Operators of Loop Quantum Gravity?, [gr-qc/0207112]
1475: 
1476: \bibitem{HolFluxRep}
1477: A.\ Okolow and J.\ Lewandowski,
1478: \newblock Diffeomorphism covariant representations of the holonomy-flux
1479:   star-algebra,
1480: \newblock {\em Class.\ Quantum Grav.} 20 (2003) 3543--3568, [gr-qc/0302059]
1481: 
1482: \bibitem{SuperSel}
1483: H.\ Sahlmann and T.\ Thiemann,
1484: \newblock On the superselection theory of the Weyl algebra for diffeomorphism
1485:   invariant quantum gauge theories, [gr-qc/0302090]
1486: 
1487: \bibitem{WeylRep}
1488: C.\ Fleischhack,
1489: \newblock Representations of the Weyl Algebra in Quantum Geometry,
1490:   [math-ph/0407006]
1491: 
1492: \bibitem{AreaVol}
1493: C.\ Rovelli and L.\ Smolin,
1494: \newblock Discreteness of Area and Volume in Quantum Gravity,
1495: \newblock {\em Nucl.\ Phys.\ B} 442 (1995) 593--619, [gr-qc/9411005],
1496: \newblock Erratum: {\em Nucl.\ Phys.\ B} 456 (1995) 753
1497: 
1498: \bibitem{Area}
1499: A.\ Ashtekar and J.\ Lewandowski,
1500: \newblock Quantum Theory of Geometry I: Area Operators,
1501: \newblock {\em Class.\ Quantum Grav.} 14 (1997) A55--A82, [gr-qc/9602046]
1502: 
1503: \bibitem{Vol2}
1504: A.\ Ashtekar and J.\ Lewandowski,
1505: \newblock Quantum Theory of Geometry II: Volume Operators,
1506: \newblock {\em Adv.\ Theor.\ Math.\ Phys.} 1 (1997) 388--429, [gr-qc/9711031]
1507: 
1508: \bibitem{QSDI}
1509: T.\ Thiemann,
1510: \newblock Quantum Spin Dynamics {(QSD)},
1511: \newblock {\em Class.\ Quantum Grav.} 15 (1998) 839--873, [gr-qc/9606089]
1512: 
1513: \bibitem{QSDV}
1514: T.\ Thiemann,
1515: \newblock {QSD V}: Quantum Gravity as the Natural Regulator of Matter Quantum
1516:   Field Theories,
1517: \newblock {\em Class.\ Quantum Grav.} 15 (1998) 1281--1314, [gr-qc/9705019]
1518: 
1519: \bibitem{PhD}
1520: M.\ Bojowald,
1521: \newblock {\em Quantum Geometry and Symmetry},
1522: \newblock PhD thesis, RWTH Aachen, 2000,
1523: \newblock published by Shaker-Verlag, Aachen
1524: 
1525: \bibitem{SymmRed}
1526: M.\ Bojowald and H.~A.\ Kastrup,
1527: \newblock Symmetry Reduction for Quantized Diffeomorphism Invariant Theories of
1528:   Connections,
1529: \newblock {\em Class.\ Quantum Grav.} 17 (2000) 3009--3043, [hep-th/9907042]
1530: 
1531: \bibitem{SphSymm}
1532: M.\ Bojowald,
1533: \newblock Spherically Symmetric Quantum Geometry: States and Basic Operators,
1534: \newblock {\em Class.\ Quantum Grav.} 21 (2004) 3733--3753, [gr-qc/0407017]
1535: 
1536: \bibitem{Bohr}
1537: A.\ Ashtekar, M.\ Bojowald, and J.\ Lewandowski,
1538: \newblock Mathematical structure of loop quantum cosmology,
1539: \newblock {\em Adv.\ Theor.\ Math.\ Phys.} 7 (2003) 233--268, [gr-qc/0304074]
1540: 
1541: \bibitem{PolymerParticle}
1542: A.\ Ashtekar, S.\ Fairhurst, and J.\ Willis,
1543: \newblock Quantum Gravity, Shadow States, and Quantum Mechanics,
1544: \newblock {\em Class.\ Quant.\ Grav.} 20 (2003) 1031--1062, [gr-qc/0207106]
1545: 
1546: \bibitem{QEDBohr}
1547: W.\ Thirring and H.\ Narnhofer,
1548: \newblock Covariant QED without Infinite Metrics,
1549: \newblock {\em Rev.\ Math.\ Phys.} SI1 (1992) 197--211
1550: 
1551: \bibitem{LoopCosRev} 
1552: M.\ Bojowald and H.~A.\ Morales-T\'ecotl,
1553:   \newblock Cosmological applications of loop quantum gravity,
1554:   \newblock In {\em Proceedings of the Fifth Mexican School (DGFM):
1555:     The Early Universe and Observational Cosmology},
1556:   {\em Lect.\ Notes Phys.} {\bf 646} (2004), 421--462
1557:   (Springer-Verlag), [gr-qc/0306008]
1558: 
1559: \bibitem{IsoCosmo}
1560: M.\ Bojowald,
1561: \newblock Isotropic Loop Quantum Cosmology,
1562: \newblock {\em Class.\ Quantum Grav.} 19 (2002) 2717--2741, [gr-qc/0202077]
1563: 
1564: \bibitem{SemiClass}
1565: M.\ Bojowald,
1566: \newblock The Semiclassical Limit of Loop Quantum Cosmology,
1567: \newblock {\em Class.\ Quantum Grav.} 18 (2001) L109--L116, [gr-qc/0105113]
1568: 
1569: \bibitem{Decoherence}
1570: D.\ Giulini, C.\ Kiefer, E.\ Joos, J.\ Kupsch, I.~O.\ Stamatescu, and H.~D.\
1571:   Zeh,
1572: \newblock {\em Decoherence and the Appearance of a Classical World in Quantum
1573:   Theory},
1574: \newblock Springer, Berlin, Germany, 1996
1575: 
1576: \bibitem{Golam}
1577: G.~M.\ Hossain,
1578: \newblock Hubble operator in isotropic loop quantum cosmology,
1579: \newblock {\em Class.\ Quantum Grav.} 21 (2004) 179--196, [gr-qc/0308014]
1580: 
1581: \bibitem{IsoSpinFoam}
1582: K.\ Noui, A.\ Perez, and K.\ Vandersloot,
1583: \newblock On the Physical Hilbert Space of Loop Quantum Cosmology,
1584: \newblock {\em Phys.\ Rev.\ D} 71 (2005) 044025, [gr-qc/0411039]
1585: 
1586: \bibitem{FundamentalDisc}
1587: M.\ Bojowald and G.\ Date,
1588: \newblock Consistency conditions for fundamentally discrete theories,
1589: \newblock {\em Class.\ Quantum Grav.} 21 (2004) 121--143, [gr-qc/0307083]
1590: 
1591: \bibitem{DynIn}
1592: M.\ Bojowald,
1593: \newblock Dynamical Initial Conditions in Quantum Cosmology,
1594: \newblock {\em Phys.\ Rev.\ Lett.} 87 (2001) 121301, [gr-qc/0104072]
1595: 
1596: \bibitem{Essay}
1597: M.\ Bojowald,
1598: \newblock Initial Conditions for a Universe,
1599: \newblock {\em Gen.\ Rel.\ Grav.} 35 (2003) 1877--1883, [gr-qc/0305069]
1600: 
1601: \bibitem{GenFunc}
1602: D.\ Cartin, G.\ Khanna, and M.\ Bojowald,
1603: \newblock Generating function techniques for loop quantum cosmology,
1604: \newblock {\em Class.\ Quantum Grav.} 21 (2004) 4495--4509, [gr-qc/0405126]
1605: 
1606: \bibitem{GenFuncBI}
1607: D.\ Cartin and G.\ Khanna,
1608: \newblock Absence of pre-classical solutions in Bianchi I loop quantum
1609:   cosmology,
1610: \newblock {\em Phys.\ Rev.\ Lett.} 94 (2005) 111302, [gr-qc/0501016]
1611: 
1612: \bibitem{InvScale}
1613: M.\ Bojowald,
1614: \newblock Inverse Scale Factor in Isotropic Quantum Geometry,
1615: \newblock {\em Phys.\ Rev.\ D} 64 (2001) 084018, [gr-qc/0105067]
1616: 
1617: \bibitem{Gaul}
1618: M.\ Gaul and C.\ Rovelli,
1619: \newblock A generalized Hamiltonian Constraint Operator in Loop Quantum Gravity
1620:   and its simplest Euclidean Matrix Elements,
1621: \newblock {\em Class.\ Quantum Grav.} 18 (2001) 1593--1624, [gr-qc/0011106]
1622: 
1623: \bibitem{Ambig}
1624: M.\ Bojowald,
1625: \newblock Quantization ambiguities in isotropic quantum geometry,
1626: \newblock {\em Class.\ Quantum Grav.} 19 (2002) 5113--5130, [gr-qc/0206053]
1627: 
1628: \bibitem{ICGC}
1629: M.\ Bojowald,
1630: \newblock Loop Quantum Cosmology: Recent Progress,
1631: \newblock In {\em Proceedings of the International Conference on Gravitation
1632:   and Cosmology (ICGC 2004), Cochin, India},
1633:   {\em Pramana} {\bf 63} (2004) 765--776, [gr-qc/0402053]
1634: 
1635: \bibitem{Perturb}
1636: A.\ Ashtekar, M.\ Bojowald, and J.\ Willis, in preparation
1637: 
1638: \bibitem{Josh}
1639: J.\ Willis,
1640: \newblock {\em On the Low-Energy Ramifications and a Mathematical Extension of
1641:   Loop Quantum Gravity},
1642: \newblock PhD thesis, The Pennsylvania State University, 2004
1643: 
1644: \bibitem{Time}
1645: M.\ Bojowald, P.\ Singh, and A.\ Skirzewski,
1646: \newblock Time dependence in quantum gravity,
1647: \newblock {\em Phys.\ Rev.\ D} 70 (2004) 124022, [gr-qc/0408094]
1648: 
1649: \bibitem{BounceClosed}
1650: P.\ Singh and A.\ Toporensky,
1651: \newblock Big Crunch Avoidance in ${\rm k}=1$ Semi-Classical Loop Quantum
1652:   Cosmology,
1653: \newblock {\em Phys.\ Rev.\ D} 69 (2004) 104008, [gr-qc/0312110]
1654: 
1655: \bibitem{BounceQualitative}
1656: G.~V.\ Vereshchagin,
1657: \newblock Qualitative Approach to Semi-Classical Loop Quantum Cosmology,
1658: \newblock {\em JCAP} 07 (2004) 013, [gr-qc/0406108]
1659: 
1660: \bibitem{Oscill}
1661: J.~E.\ Lidsey, D.~J.\ Mulryne, N.~J.\ Nunes, and R.\ Tavakol,
1662: \newblock Oscillatory Universes in Loop Quantum Cosmology and Initial
1663:   Conditions for Inflation,
1664: \newblock {\em Phys.\ Rev.\ D} 70 (2004) 063521, [gr-qc/0406042]
1665: 
1666: \bibitem{Cyclic}
1667: M.\ Bojowald, R.\ Maartens, and P.\ Singh,
1668: \newblock Loop Quantum Gravity and the Cyclic Universe,
1669: \newblock {\em Phys.\ Rev.\ D} 70 (2004) 083517, [hep-th/0407115]
1670: 
1671: \bibitem{EffHam}
1672: G.\ Date and G.~M.\ Hossain,
1673: \newblock Effective Hamiltonian for Isotropic Loop Quantum Cosmology,
1674: \newblock {\em Class.\ Quantum Grav.} 21 (2004) 4941--4953, [gr-qc/0407073]
1675: 
1676: \bibitem{GenericBounce}
1677: G.\ Date and G.~M.\ Hossain,
1678: \newblock Genericity of Big Bounce in isotropic loop quantum cosmology,
1679: \newblock {\em Phys.\ Rev.\ Lett.} 94 (2005) 011302, [gr-qc/0407074]
1680: 
1681: \bibitem{GenericInfl}
1682: G.\ Date and G.~M.\ Hossain,
1683: \newblock Genericity of inflation in isotropic loop quantum cosmology,
1684: \newblock {\em Phys.\ Rev.\ Lett.} 94 (2005) 011301, [gr-qc/0407069]
1685: 
1686: \bibitem{Horizon}
1687: M.\ Bojowald and R.\ Swiderski,
1688: \newblock Spherically Symmetric Quantum Horizons,
1689: \newblock {\em Phys.\ Rev.\ D} 71 (2005) 081501(R), [gr-qc/0410147]
1690: 
1691: \bibitem{SphSymmHam}
1692: M.\ Bojowald and R.\ Swiderski,
1693: \newblock Spherically Symmetric Quantum Geometry: Hamiltonian Constraint,
1694: \newblock to appear
1695: 
1696: \bibitem{PowerLoop}
1697: G.~M.\ Hossain,
1698: \newblock Primordial Density Perturbation in Effective Loop Quantum Cosmology,
1699: \newblock {\em Class.\ Quantum Grav.} 22 (2005) (to appear), [gr-qc/0411012]
1700: 
1701: \bibitem{InflationWMAP}
1702: S.\ Tsujikawa, P.\ Singh, and R.\ Maartens,
1703: \newblock Loop quantum gravity effects on inflation and the CMB,
1704: \newblock {\em Class.\ Quantum Grav.} 21 (2004) 5767--5775, [astro-ph/0311015]
1705: 
1706: \bibitem{Robust}
1707: M.\ Bojowald, J.~E.\ Lidsey, D.~J.\ Mulryne, P.\ Singh, and R.\ Tavakol,
1708: \newblock Inflationary Cosmology and Quantization Ambiguities in Semi-Classical
1709:   Loop Quantum Gravity,
1710: \newblock {\em Phys.\ Rev.\ D} 70 (2004) 043530, [gr-qc/0403106]
1711: 
1712: \bibitem{InflOsc}
1713: D.~J.\ Mulryne, N.~J.\ Nunes, R.\ Tavakol, and J.\ Lidsey,
1714: \newblock Inflationary Cosmology and Oscillating Universes in Loop Quantum
1715:   Cosmology,
1716: \newblock {\em Int.\ J.\ Mod.\ Phys.\ A} (2004) to appear, [gr-qc/0411125]
1717: 
1718: \bibitem{Emergent}
1719: G.~F.~R.\ Ellis and R.\ Maartens,
1720: \newblock The Emergent Universe: inflationary cosmology with no singularity,
1721: \newblock {\em Class.\ Quant.\ Grav.} 21 (2004) 223--232, [gr-qc/0211082]
1722: 
1723: \bibitem{Emergent2}
1724: G.~F.~R.\ Ellis, J.\ Murugan, and C.~G.\ Tsagas,
1725: \newblock The Emergent Universe: An Explicit Construction,
1726: \newblock {\em Class.\ Quant.\ Grav.} 21 (2004) 233--250, [gr-qc/0307112]
1727: 
1728: \bibitem{EmergentLoop}
1729: D.~J.\ Mulryne, R.\ Tavakol, J.~E.\ Lidsey, and G.~F.~R.\ Ellis,
1730: \newblock An emergent universe from a loop, [astro-ph/0502589]
1731: 
1732: \bibitem{PowerPert}
1733: S.\ Hofmann and O.\ Winkler,
1734: \newblock The Spectrum of Fluctuations in Inflationary Quantum Cosmology,
1735:   [astro-ph/0411124]
1736: 
1737: \bibitem{HomCosmo}
1738: M.\ Bojowald,
1739: \newblock Homogeneous loop quantum cosmology,
1740: \newblock {\em Class.\ Quantum Grav.} 20 (2003) 2595--2615, [gr-qc/0303073]
1741: 
1742: \bibitem{Spin}
1743: M.\ Bojowald, G.\ Date, and K.\ Vandersloot,
1744: \newblock Homogeneous loop quantum cosmology: The role of the spin connection,
1745: \newblock {\em Class.\ Quantum Grav.} 21 (2004) 1253--1278, [gr-qc/0311004]
1746: 
1747: \bibitem{BoundFull}
1748: J.\ Brunnemann and T.\ Thiemann,
1749: \newblock Unboundedness of Triad-Like Operators in Loop Quantum Gravity,
1750:   [gr-qc/0505033]
1751: 
1752: \bibitem{BKL}
1753: V.~A.\ Belinskii, I.~M.\ Khalatnikov, and E.~M.\ Lifschitz,
1754: \newblock A general solution of the Einstein equations with a time singularity,
1755: \newblock {\em Adv.\ Phys.} 13 (1982) 639--667
1756: 
1757: \bibitem{NonChaos}
1758: M.\ Bojowald and G.\ Date,
1759: \newblock Quantum suppression of the generic chaotic behavior close to
1760:   cosmological singularities,
1761: \newblock {\em Phys.\ Rev.\ Lett.} 92 (2004) 071302, [gr-qc/0311003]
1762: 
1763: \bibitem{ChaosLQC}
1764: M.\ Bojowald, G.\ Date, and G.~M.\ Hossain,
1765: \newblock The Bianchi IX model in loop quantum cosmology,
1766: \newblock {\em Class.\ Quantum Grav.} 21 (2004) 3541--3569, [gr-qc/0404039]
1767: 
1768: \bibitem{Collapse}
1769: M.\ Bojowald, R.\ Goswami, R.\ Maartens, and P.\ Singh,
1770: \newblock A black hole mass threshold from non-singular quantum gravitational
1771:   collapse, [gr-qc/0503041]
1772: 
1773: \bibitem{Brownian}
1774: A.\ Einstein,
1775: \newblock \"Uber die von der molekularkinetischen Theorie der W\"arme
1776:   geforderte Bewegung von in ruhenden Fl\"ussigkeiten suspendierten Teilchen,
1777: \newblock {\em Annalen Phys.} 17 (1905) 549--560
1778: 
1779: \end{thebibliography}
1780: 
1781: 
1782: \end{document}