1: \documentclass[preprint,12pt,tightenlines,eqsecnum,floats,aps,amsmath,amssymb%
2: ,nofootinbib,prd,showpacs]{revtex4}
3:
4: \usepackage{amsmath,amssymb,amsfonts}
5: \usepackage{graphicx}
6: \usepackage{enumerate}
7: %\usepackage{colordvi} % for color text
8:
9: \usepackage{LQC-symbols}
10:
11:
12: \begin{document}
13:
14: \preprint{\vbox{\baselineskip=12pt \rightline{IGC-08/3-3}
15: }}
16:
17: \title{Anti-deSitter universe dynamics in LQC}
18:
19: \author{Eloisa Bentivegna${}^{2,3}$}
20: \email{bentiveg@gravity.psu.edu}
21: \author{Tomasz Pawlowski${}^{1,2}$}
22: \email{tomasz@iem.cfmac.csic.es}
23: \affiliation{
24: ${}^{1}$Instituto de Estructura de la Materia, \\
25: Consejo Superior de Investigaciones Cient\'ificas (CSIC), \\
26: Serrano 121, 28006 Madrid, Spain\\
27: ${}^{2}$Institute for Gravitational Physics and Geometry,\\
28: Physics Department, Penn State, \\
29: University Park, PA 16802, U.S.A.\\
30: ${}^{3}$Center for Gravitational Wave Physics,\\
31: Physics Department, Penn State, \\
32: University Park, PA 16802, U.S.A.
33: }
34:
35: \begin{abstract}
36: A model for a flat isotropic universe with a negative cosmological
37: constant $\Lambda$ and a massless scalar field as sole matter content is
38: studied within the framework of Loop Quantum Cosmology. By application
39: of the methods introduced for the model with $\Lambda=0$, the physical
40: Hilbert space and the set of Dirac observables are constructed. As in
41: that case, the scalar field plays here the role of an emergent
42: time. The properties of the system are found to be similar to those of the
43: $k=1$ FRW model: for small energy densities, the quantum dynamics reproduces
44: the classical one, whereas, due to modifications at near-Planckian
45: densities, the big bang and big crunch singularities are replaced by a
46: quantum bounce connecting deterministically the large semiclassical
47: epochs. Thus in Loop Quantum Cosmology the evolution is
48: qualitatively cyclic.
49: \end{abstract}
50:
51:
52: \pacs{04.60.Kz, 04.60.Pp, 98.80.Qc, 03.65.Sq}
53:
54: \maketitle
55:
56:
57: \section{Introduction}
58: \label{sec:intro}
59:
60: Loop Quantum Cosmology \cite{lqc} -- an application of methods of Loop
61: Quantum Gravity \cite{lqg} to symmetry reduced models -- constitutes a
62: promising way of studying quantum-gravitational effects in
63: cosmological models. In particular one of the simplest models, a flat
64: Friedmann-Robertson-Walker (FRW) universe was analyzed within its
65: framework \cite{aps,aps-old,aps-imp}. In that case, the structure of the
66: Hamiltonian constraint allowed to treat the constrained system as a
67: free one, evolving with respect to the scalar field which thus plays the
68: role of an emergent time. This, in turn, allowed the construction of
69: a physical Hilbert space and a set of Dirac observables, which were used
70: next to extract the physics by means of numerical methods. The results were
71: quite surprising: the analysis has shown that, when the matter energy
72: density approaches the Planck scale, the quantum-geometric effects
73: cause gravity to become repulsive. In consequence, a large
74: semiclassical expanding universe is preceded by a (also large and
75: semiclassical) contracting one, deterministically connected to the
76: former by a quantum bridge. The transition point of the evolution
77: (called quantum bounce) is characterized by an energy density which, at
78: this point, equals the critical value $\rho_c \approx 0.82\rhoPl$.
79: Furthermore, even when quantum corrections actually dominate the
80: dynamics, the state representing the universe remains semiclassical --
81: its evolution is to great precision described by the so called
82: classical effective dynamics \cite{pv,aps-imp}.
83:
84: The results obtained for the flat FRW model were next generalized to
85: the spherical one \cite{apsv} (the $k=1$ FRW model). The properties of
86: the Hilbert space and an evolution operator were investigated analytically
87: \cite{klp,klp-nL} and the robustness of their features was confirmed
88: through the analysis of its approximation (known as {\it sLQC})
89: \cite{acs,cs-rec}. Further generalizations to anisotropic (and
90: further inhomogeneous) models by different research groups are in
91: various stages of progress \cite{chv,gp,gm-letter}.
92:
93: Thus far, however, the only models described rigorously were universes
94: with a vanishing cosmological constant $\Lambda$ and a massless
95: scalar field. In this article, we extend the analysis of \cite{aps-imp}
96: to include the universes with negative $\Lambda$.
97: Although the observations favor a positive $\Lambda$, this model
98: constitutes a convenient way of testing which features of the previously
99: investigated model we can hope to generalize to more realistic
100: systems. Also, since it is a classically recollapsing system, we can use
101: it to investigate semiclassicality issues (dispersion after many 'cycles'
102: of evolution). The specific questions we intend to address here are the
103: following:
104: \begin{itemize}
105: \item Do the qualitative features of the $\Lambda=0$ model survive
106: also in this case? In particular, are the big bang/crunch
107: singularities replaced by quantum bounces as in the previously
108: investigated cases? All the models analyzed so far not only
109: experienced the bounce, but for Gaussian states the observed
110: dispersion of the wave packet after the bounce was severely restricted
111: by the values of the spreads before it. In the flat case this result
112: was next generalized analytically to a space of states admitting
113: semiclassical epoch\footnote{The states for which either
114: at early or late times the relative dispersions of chosen Dirac
115: observables are $\ll 1$.} \cite{cs-rec} within the context of
116: sLQC. Therefore,
117: it is important to ask whether such behavior will occur also in
118: the considered model, or it was just a result of the extreme
119: simplicity of the previous ones.
120: \item If the answer to the previous question is in the affirmative, then
121: is the critical energy density $\rho_c$ still a fundamental bound?
122: In both the $k=0$ and $k=1$ models for physically
123: sensible\footnote{This indicates the states of the scalar field
124: with momentum sufficiently high for the closed universe to grow
125: to macroscopic ($>1$ megaparsec) scales before recollapsing.}
126: states, the matter energy density at the bounce point agreed to
127: great precision with $\rho_c$. Furthermore, later investigations
128: within the sLQC model have shown that $\rho_c$ is indeed a
129: fundamental energy bound. But again, we do not know a priori
130: whether this feature is characteristic just of the models
131: investigated so far and how (if at all) it generalizes.
132: \item Does this model possess any new feature not observed in
133: $\Lambda=0$ or $k=1$ case ?
134: \end{itemize}
135: A preliminary investigation of the $\Lambda<0$ model has been conducted
136: already in \cite{aps-imp}. However, the physical Hilbert space was not
137: constructed; the goal there was only to verify the persistence of the
138: bounce. Recently, a heuristically constructed effective classical
139: Hamiltonian was used \cite{eff} to obtain the effective trajectories
140: of both the $\Lambda<0$ and $\Lambda>0$ systems and analyze the effect
141: of the quantum-geometric corrections on the universe's dynamics.
142: However, since the effective Hamiltonian was not derived systematically,
143: the results have to be confirmed against genuine quantum evolution.
144:
145: In addition to the problems described above, we also address the concerns
146: about the choice of the symmetric sector of the physical Hilbert space
147: that is sometimes raised. Because of the absence of fermions, the triad
148: orientation reflection is a large gauge symmetry. This allowed one to
149: restrict the physical Hilbert space to the states symmetric under parity
150: reflection. However, since the choice of antisymmetric states
151: is equally justified, it is natural to ask whether the results of LQC
152: are robust and will continue to hold if the antisymmetric sector
153: is chosen. We address this issue by analyzing, in addition to the standard
154: symmetric states, also the space of antisymmetric ones and establish
155: robustness.
156:
157: The paper is organized as follows:
158: %
159: we start with a brief summary of the basic framework (introduced
160: already in earlier papers) in section \ref{sec:frame}. Its content is
161: divided into three parts: the classical theory, the kinematics of LQC
162: and the derivation of the quantum Hamiltonian constraint.
163: %
164: In section \ref{sec:WDW} we consider a geometrodynamical equivalent of
165: the model -- the Wheeler-DeWitt (WDW) one. The reason for that is
166: two-fold: first, it will allow us to compare the results of LQC
167: against a standard quantum model and identify the nonperturbative
168: quantum-geometric effects. Second, it will serve as an introduction to
169: the methodology of extracting physics, used next on the LQC model. The
170: analytical solvability of the WDW model will allow us to
171: show these methods without having to deal with the complications of
172: numerical analysis.
173: %
174: Analysis of the physical sector is carried out in section
175: \ref{sec:LQC}. There, we extensively use the results of the numerical
176: study described in turn in section \ref{sec:num}. That section
177: contains also a description of the construction and analysis of the
178: states semiclassical at late times. The final results and their
179: discussion are placed in section \ref{sec:concl}.
180:
181: Apart from the main body, the article contains two appendices: in section
182: \ref{sec:antisymm}, we analyze the space of antisymmetric states,
183: whereas \ref{sec:heu} contains a description of the heuristic methods used
184: to extract some of the results.
185:
186:
187: \section{The LQC quantization scheme}
188: \label{sec:frame}
189:
190: In this section, we introduce the quantization framework used in later
191: sections of the paper. Since we directly apply the framework described
192: in detail in \cite{aps-imp,apsv}, we will just present a brief sketch
193: of it. For a more detailed discussion, the reader is referred to the
194: above mentioned articles.
195:
196: The content of this section is divided into three parts. In the first,
197: we present the classical theory used as a basis for quantization.
198: The second part is dedicated to the description of the LQC kinematics.
199: Finally, we recall the derivation of the LQC Hamiltonian constraint.
200:
201:
202: \subsection{Classical theory}
203: \label{sec:frameClass}
204:
205: A flat ($k=0$) FRW model represents a spacetime admitting a foliation by
206: spatial isotropic $3$-surfaces $M$ of topology $\re^3$. Its metric
207: tensor can be written in the form
208: \begin{equation}
209: g = -\rd t^2 + a^2(t) \fidq \ ,
210: \end{equation}
211: where $t$ is a time parameter (the {\it cosmic time}), $\fidq$ is a unit
212: (fiducial) Cartesian metric on the surface $M$ and the function $a(t)$ is
213: called a scale factor.
214:
215: Due to the homogeneity and noncompactness of $M$, one cannot write an
216: action or Hamiltonian as an integral of the appropriate density over the
217: entire $M$. Instead, we can define them as integrals over a chosen fiducial
218: cubical cell $\fV$, constant in comoving coordinates\footnote{The considered
219: model is of the Bianchi type A: the equations of motion derived from
220: the Hamiltonian specified in this way are identical to the Einstein
221: field equations reduced to the isotropic case.}. Given such a cell,
222: one can define a triad $\fide$ (and cotriad $\fidw$ dual to it) as
223: directed along the edges of $\fV$ and orthonormal with respect to $\fidq$.
224:
225: As gravitational phase space variables, we choose the connections
226: $A^i_a$ and the density-weighted triads $E^a_i$
227: \begin{subequations}\label{eq:AE-def}\begin{align}
228: A^i_a\ &=\ c V_o^{-\frac{1}{3}}\,\fidw^i_a \ , &
229: E^a_i\ &=\ p V_o^{-\frac{2}{3}}\sqrt{\fidq}\,\fide^a_i \ ,
230: \tag{\ref{eq:AE-def}}
231: \end{align}\end{subequations}
232: where $V_o$ is a volume of $\fV$ with respect to $\fidq$. The real
233: parameters $c,p$ called respectively {\it connection} and {\it triad
234: coefficients}
235: coordinatize the ($2$-dimensional) phase space of the gravitational
236: degrees of freedom. Appropriate scaling by $V_o$ ensures the invariance of
237: the symplectic structure of this phase space (when written in terms of
238: $c,p$) under different choices of $\fidq$. The Poisson bracket
239: between $c$ and $p$ equals
240: \begin{equation}
241: \{c,p\}\ =\ \frac{8\pi\gamma G}{3} \ ,
242: \end{equation}
243: where $\gamma$ is the Barbero-Immirzi parameter.
244:
245: The basic variables defined as in \eqref{eq:AE-def} automatically
246: satisfy the Gauss and diffeomorphism constraints. The contribution of
247: the geometry to the only nontrivial constraint -- the Hamiltonian
248: one -- is of the form
249: \begin{equation}\label{eq:classHgrav}
250: \Cgrav\ =\ -\frac{1}{\gamma^2} \int_{\fV} \rd^3 x \left( \varepsilon_{ijk}
251: e^{-1} E^{ai} E^{bj} F^k_{ab} - \gamma^2\Lambda \right)
252: \ =\ -\frac{6}{\gamma^2} c^2\sqrt{p}
253: + {\Lambda}p^{\frac{3}{2}}
254: \end{equation}
255: where $e\ :=\ \sqrt{|\det E|}$ and the field strength
256: $F^k_{ab}\ :=\ 2\partial_a A^k_b + \varepsilon^k_{ij} A^i_a A^j_b$.
257:
258: The only matter content -- a homogeneous massless scalar field -- is
259: described by two global variables: the field value $\phi$ and its
260: conjugate momentum $p_{\phi}$, with Poisson bracket between them
261: \begin{equation}
262: \{\phi,p_{\phi}\}\ =\ 1 \ .
263: \end{equation}
264: The pair $(\phi,p_{\phi})$ coordinatizes the phase space corresponding
265: to the matter degrees of freedom. The full phase space of the system
266: is thus $4$-dimensional. The complete Hamiltonian constraint is of the form
267: \begin{equation}\label{eq:classHcompl}
268: C\ =:\ \Cgrav + \Cphi\ =\ 0 \ , \qquad
269: \text{where}\ \ \Cphi\ =\ 8\pi G p^{-\frac{3}{2}} p^2_{\phi} \ .
270: \end{equation}
271: The above constraint defines a $3$D hypersurface in the $4$D phase
272: space. Furthermore, since $C$ does not depend explicitly on $\phi$, the
273: momentum $p_{\phi}$ is a constant of motion. Therefore, the dynamical
274: trajectories can be represented as a (parametrized by $p_{\phi}$) family
275: of functions $p(\phi)$
276: \begin{equation}\label{eq:pre-class-traj}
277: p(\phi)\ =\ \frac{(4\pi G)^{\frac{1}{3}}p_{\phi}^{\frac{2}{3}}}%
278: {|\Lambda|^{\frac{1}{3}}\cosh(\sqrt{12\pi G}(\phi-\phi_o))}
279: \end{equation}
280: Their form implies that the considered system recollapses. Each
281: trajectory starts at the big bang singularity and ends in a big crunch.
282:
283:
284: \subsection{Kinematics of LQC}
285: \label{sec:frameKin}
286:
287: To quantize the system, we follow the Dirac program. First we construct a
288: kinematical Hilbert space: in our case, it is the tensor product of
289: spaces corresponding to, respectively, gravitational and matter degrees
290: of freedom: $\Hilk = \Hilkg\otimes\Hilkf$.
291:
292: For the matter we apply the standard Schr{\"o}dinger quantization.
293: As $\Hilkf$ we choose the standard Hilbert space of square integrable
294: functions $\Hilkf = L^2(\re,\rd\phi)$. The basic operators are
295: $\hat{\phi}$ and $\hat{p}_{\phi}$. To describe the state we choose the
296: (dual) basis $\dbra{\phi}$ of eigenstates of $\hat{\phi}$. The action
297: of $\hat{\phi}$, $\hat{p}_{\phi}$ on the state can be then expressed
298: as follows
299: \begin{subequations}\label{eq:phiOp}\begin{align}
300: \hat{\phi}\Psi(\phi)\ &=\ \phi\Psi(\phi) \ , &
301: \hat{p}_{\phi}\Psi(\phi)\ &=\ -i\hbar\partial_{\phi}\Psi(\phi) \ , &
302: \text{where}\ \Psi(\phi)\ :=\ \dip{\phi}{\Psi} \ .
303: \tag{\ref{eq:phiOp}}\end{align}\end{subequations}
304:
305: The quantization of the gravitational degrees of freedom within LQC at the
306: kinematical level has been rigorously performed in \cite{abl}. The
307: procedure is the analog of the quantization scheme used in full LQG
308: (see for example \cite{lqc-MB}).
309: Here the basic variables are triads and connections along straight
310: edges generated by $\fide^a_i$. The kinematical Hilbert space is
311: the space of square integrable functions on the Bohr compactification of
312: the real line $\Hilkg = L^2(\rBohr,\rd\mu_{\Bohr})$. We will represent
313: its elements using the basis consisting of the eigenfunctions of $p$
314: (promoted to an operator), labeled by $\mu\in\re$. Despite
315: the continuity of $\mu$, the elements of the chosen basis are
316: orthonormal with respect to Kronecker delta
317: \begin{equation}\label{eq:kinIP}
318: \sip{\mu_1}{\mu_2}\ =\ \delta_{\mu_1\mu_2} \ .
319: \end{equation}
320: As basic quantum operators, we select $\hat{p}$ and $\whexp(i\fracs{\lambda
321: c}{2})$ \footnote{Since the family $\whexp(i\lambda c/2)$ is not
322: weakly continuous, the operator $\hat{c}$ does not exists.}. Their
323: action on the basis elements $\sket{\mu}$ is given by:
324: \begin{subequations}\label{eq:ph-act}\begin{align}
325: \hat{p}\sket{\mu}\ &=\ \frac{8\pi\gamma G\lPl^2}{6}\sket{\mu} \ , &
326: \whexp(i\fracs{\lambda c}{2})\sket{\mu}\ &=\ \sket{\mu+\lambda}
327: \ .
328: \tag{\ref{eq:ph-act}}
329: \end{align}\end{subequations}
330: Since the holonomy along the edge of fiducial length $\lambda$
331: generated by $\fide^a_i$ can be expressed via $\exp(i\lambda c/2)$
332: \begin{equation}\label{eq:hexp}
333: \hol{\lambda}{k}\ =\ \fracs{1}{2}[\exp(\fracs{i\lambda c}{2})
334: + \exp(-\fracs{i\lambda c}{2})]\id
335: +\fracs{1}{i} [\exp(\fracs{i\lambda c}{2})
336: - \exp(-\fracs{i\lambda c}{2})] \tau_k
337: \end{equation}
338: (where the $\tau_k$ are related to the Pauli matrices $\sigma_k$ via
339: $2i\tau_k=\sigma_k$), its quantum analog $\qhol{\lambda}{k}$ can
340: be expressed in terms of the operators $\whexp$ in the same way.
341:
342:
343: \subsection{LQC: the Hamiltonian constraint}
344: \label{sec:frameH}
345:
346: In order to write the quantum operator corresponding to the Hamiltonian
347: constraint (\ref{eq:classHgrav},~\ref{eq:classHcompl}), we need to
348: reexpress it in terms of the basic objects selected in the previous
349: subsection.
350:
351: Let us start with $\Cgrav$ \eqref{eq:classHgrav}. The quantization of the
352: cosmological term is straightforward (and just amounts to promoting
353: $p$ to operator $\hat{p}$). The remaining part is an integral of the
354: product of two terms: $e^{-1} E^{ai} E^{bj}$ and $F^k_{ab}$.
355:
356: Following Thiemann \cite{ThTrick}, we can rewrite the first term
357: in the following form
358: \begin{equation}\label{eq:eEE}
359: \varepsilon_{ijk}e^{-1}E^{ai}E^{bj}\ =\ \sum_k
360: \frac{\sgn(p)}{2\pi\gamma G\lambda V_o^{\frac{1}{3}}}
361: \fidveps^{abc} \fidw^k_c \Tr \left( h^{(\lambda)}_k
362: \{{h^{(\lambda)}_k}^{-1},V\}\tau_i \right)
363: \end{equation}
364: where $V=|p|^{\frac{3}{2}}$ is the (physical) volume of the cell $\fV$.
365:
366: The field strength term $F^k_{ab}$ can, on the other hand, be approximated
367: via holonomies along the square loop $\square_{ij}$ oriented on the
368: $i$-$j$ plane.
369: \begin{subequations}\label{eq:Fdef}\begin{align}
370: F^k_{ab}\ &=\ -2
371: \Tr\left( \frac{\hol{\lambda}{\square_{ij}}-1}{\lambda^2
372: V_o^{\frac{2}{3}}} \right)
373: \tau^k \fidw^i_a \fidw^j_b \ , &
374: \hol{\lambda}{\square_{ij}}\ &=\
375: \hol{\lambda}{i} \hol{\lambda}{j} \hol{\lambda}{i}^{-1}
376: \hol{\lambda}{j}^{-1} \ .
377: \tag{\ref{eq:Fdef}}
378: \end{align}\end{subequations}
379: The size of $\square_{ij}$ is fixed by the requirement that its
380: physical area equals the lowest nonzero eigenvalue of the LQG area
381: operator
382: \begin{equation}\label{eq:mubar}
383: \lambda\ =\ \bar{\mu}(\mu)\ \ \ \text{s.t.}\ \ \
384: \Ar{}_{\square_{ij}}\ =\ \bar{\mu}^2|p|\ =\ \Delta\
385: :=\ (2\sqrt{3}\pi\gamma)\lPl^2 \ .
386: \end{equation}
387: To express the action of the operator corresponding to $\hol{\bar{\mu}}{}$,
388: it is convenient to use, instead of the label $\mu$, a new label $v$ defined
389: as follows
390: \begin{equation}
391: v\ :=\ K\sgn(\mu)|\mu|^{\frac{3}{2}} \ ,
392: \qquad K\ :=\ \frac{2\sqrt{2}}{3\sqrt{3\sqrt{3}}} \ .
393: \end{equation}
394: In the new labeling an exponent operator
395: $\widehat{\exp}(\frac{i\bar{\mu}c}{2})$ --the component of
396: $\hol{\bar{\mu}}{}$ (via \eqref{eq:hexp})-- acts simply as a unit
397: translation
398: \begin{equation}
399: \widehat{\exp} ( \fracs{i}{2}{\bar{\mu}c} ) \sket{v}\ =\ \sket{v+1}
400: \ .
401: \end{equation}
402:
403: In the matter part of the Hamiltonian constraint, the only nontrivial
404: component is $|p|^{-3/2}$, but again this can be reexpressed in terms of
405: holonomies via Thiemann's method
406: \begin{equation}\label{eq:invVdef}
407: |p|^{-\frac{3}{2}}\ =\ \sgn(p)\left[ \frac{1}{2\pi\lPl^2\gamma\bar{\mu}}
408: \Tr\sum_k \tau^k h^{(\bar{\mu})}_k
409: \{ {h^{(\bar{\mu})}_k}^{-1} , V^{\frac{1}{3}}\}
410: \right]^3 \ .
411: \end{equation}
412:
413: Finally, applying all the results (\ref{eq:hexp}-\ref{eq:invVdef}) to
414: \eqref{eq:classHcompl}, one can write the operator $\hat{C}$. We do so
415: choosing, in the process, a particular factor ordering (the so called
416: Kaminski ordering) \cite{aps-imp}, in which $\hCgrav$ is manifestly
417: symmetric and positive-definite. The action of the final result on
418: the state $\Psi\in\Hilk$ can be written in the following form
419: \begin{equation}\label{eq:Theta}
420: \partial^2_{\phi}\Psi(v,\phi)\ =\ -\Theta\Psi(v,\phi)\
421: =\ -\Theta_o\Psi(v,\phi) + [B(v)]^{-1} C_{\Lambda} \Psi(v,\phi)
422: \ ,
423: \end{equation}
424: where $\Psi(v,\phi):=\sip{v,\phi}{\Psi}$ and the functions $B(v)$,
425: $C_{\Lambda}(v)$ equal
426: \begin{subequations}\label{eq:BCLcoeff}\begin{align}
427: B(v)\ &:=\ \frac{27K}{8}|v|
428: \left||v+1|^{\frac{1}{3}}-|v-1|^{\frac{1}{3}}\right|^3
429: \ , &
430: C_{\Lambda}(v)\ &:=\ \frac{16\pi^2\gamma^3\lPl^4}{27K\hbar} \Lambda
431: |v|
432: \tag{\ref{eq:BCLcoeff}}
433: \end{align}\end{subequations}
434: and $\Theta_o$ is an operator corresponding to the $\Lambda=0$ case
435: derived in \cite{aps-imp}
436: \begin{equation}\label{Theta0}
437: \Theta_o\Psi(v,\phi)\
438: =\ -[B(v)]^{-1}\left( C^+(v)\Psi(v+4,\phi) + C^o(v)\Psi(v,\phi)
439: + C^-(v)\Psi(v-4,\phi) \right) \ ,
440: \end{equation}
441: with coefficients $C^{\pm}$, $C^o$, equal to
442: \begin{subequations}\label{eq:ThCoeffdef}\begin{align}
443: C^+(v)\ &=\ \frac{3\pi KG}{8}|v+2|\, \big||v+3|-|v+1|\big| \ , & & \\
444: C^-(v)\ &=\ C^+(v-4) \ ,
445: &
446: C^o(v)\ &=\ -C^+(v)-C^-(v) \ .
447: \end{align}\end{subequations}
448: For reasons we will explain in later sections of the paper, the operator
449: $\Theta$ is called an {\it evolution operator}. It is symmetric and
450: positive-definite (with respect to the measure $B(v)\rd\mu_{\Bohr}$) on
451: the domain $\Dom$ of finite linear combination of states $\sket{v}$.
452:
453:
454:
455: \section{The Wheeler-DeWitt limit}
456: \label{sec:WDW}
457:
458: The quantization scheme presented in the previous section is motivated by
459: LQG; however, it is not the only method applicable to the system.
460: By replacing $\Hilkg$ with $\ubHilkg:=L^2(\re,\rd\mu)$ and taking the
461: limit $\Delta\to 0$ in expressions
462: (\ref{eq:eEE},~\ref{eq:Fdef},~\ref{eq:invVdef}),
463: one arrives to the system equivalent to the one originating from
464: geometrodynamics, known as the Wheeler-DeWitt system. In the literature, the
465: system obtained from LQC via this procedure is called a {\it WDW
466: limit}. We will study it in this section in order to identify the
467: effects of the spacetime discreteness. We will keep this terminology
468: in the paper although (as it was shown in \cite{acs}) the WDW model
469: is not the limit of the LQC model in any precise sense. One
470: should think about it as the {\it WDW equivalent} of an LQC model.
471:
472:
473: \subsection{WDW constraint equation, emergent time}
474:
475: The evolution operator $\Theta$ is a sum of two terms: a $\Lambda=0$
476: operator $\Theta_o$ and a $\Lambda$-dependent potential term \eqref{eq:Theta}.
477: The WDW limit of $\Theta_o$ was derived in \cite{aps-imp} and is of
478: the form
479: \begin{equation}
480: \ub{\Theta}_o\,\ub{\Psi}(v,\phi)\
481: = \ -12\pi G(v\partial_v)^2\ub{\Psi}(v,\phi)
482: \end{equation}
483: where $\ub{\Psi}\in\ubHilk:=\ubHilkg\otimes\Hilkf$. Calculating the
484: limit of the cosmological constant term requires just replacing $B$ in
485: the potential term by its point limit $\ub{B} := K/|v|$ for $\Delta\to 0$.
486: In consequence, the WDW equivalent of equation \eqref{eq:Theta} has
487: the form
488: \begin{equation}\label{eq:WDWmain}
489: \partial_{\phi}^2\,\ub{\Psi}(v,\phi)\
490: =\ -\ub\Theta\,\ub{\Psi}(v,\phi)\
491: = 12\pi G(v\partial_v)^2\ub{\Psi}(v,\phi)
492: +\frac{16\pi^2\gamma^3\lPl^4}{27K^2\hbar}\Lambda
493: v^2 \ub{\Psi}(v,\phi) \ ,
494: \end{equation}
495: where the operator $\ub{\Theta}$ is symmetric and positive-definite
496: with respect to the measure $\ub{B}\rd v$ in the standard domain of
497: fast-decaying functions (Schwartz space).
498:
499: % orientation independance
500: The above constraint divides the domain of $v$ into two independent
501: sectors, corresponding to different signs of $v$, i.e. to different
502: orientations of the triad $E^a_i$. Due to the absence of a parity
503: violating interaction in the considered system, we can restrict the
504: studies to states that are symmetric/antisymmetric with respect to a reflection
505: in $v$. For further analysis, we choose the symmetric sector, that is
506: $\ub{\Psi}(\phi,v)=\ub{\Psi}(\phi,-v)$; however, the presented
507: construction can be repeated directly also in the antisymmetric case, with
508: equivalent results.
509:
510: \subsection{General solutions, frequency decomposition}
511: \label{sec:WDWsols}
512:
513: % emergent time
514: The constraint \eqref{eq:WDWmain} is similar in its form to the
515: Klein-Gordon equation. Furthermore, since there is no explicit
516: dependence on $\phi$ in either \eqref{eq:classHcompl} or
517: \eqref{eq:WDWmain}, $p_{\phi}$ is a constant of motion of both
518: the classical and the quantum system. Also, at the classical level $\phi$ is
519: monotonic in time: we can thus follow the prescription of
520: \cite{aps-imp} and reinterpret the constraint, treating it as an
521: evolution equation of a free system evolving with respect to
522: $\phi$. The scalar field becomes then an emergent time as in the case
523: $\Lambda=0$.
524:
525: % spectrum
526: To construct the physical Hilbert space we need to find the spectrum
527: of the self-adjoint extension of $\ub{\Theta}$. The eigenfunction
528: corresponding to an eigenvalue $\omega^2$ satisfying
529: \begin{equation}
530: \omega^2\ub{\psi}(v)\ =\ -12\pi G(v\partial_v)^2\ub{\psi}(v)
531: -\frac{16\pi^2\gamma^3\lPl^4}{27K^2\hbar}\Lambda
532: v^2 \ub{\psi}(v)
533: \end{equation}
534: can be written in terms of Bessel functions of the third kind
535: \begin{equation}\label{eq:WDWgeneig}
536: \psi_{\omega}(v)\ =\ c_{(I)} \BesI_{ik}(\beta\sqrt{-\Lambda}|v|)
537: + c_{(K)} \BesK_{ik}(\beta\sqrt{-\Lambda}|v|)\ ,
538: \end{equation}
539: where $k := \omega/\sqrt{12\pi G}$, $\beta :=
540: 2\sqrt{\pi\gamma^3}\hbar\lPl/(9K)$ and $c_{(I)},c_{(K)} \in \compl$.
541: When $\beta\sqrt{-\Lambda}|v|<k$, both $\BesI$ and $\BesK$ show
542: oscillatory behavior. In particular, as $|v|\to 0$, they approach
543: the eigenfunctions of the $\ub{\Theta}_o$ operator corresponding to the same
544: frequency $\omega$
545: \begin{equation}\label{eq:WDWelimit}
546: \psi_{\omega}(v)\ =\ \tilde{c}^{+}\exp(ik\ln|v|)
547: + \tilde{c}^{-}\exp(-ik\ln|v|) \ .
548: \end{equation}
549: The complex coefficients $\tilde{c}^{+}$, $\tilde{c}^{-}$ of the limit
550: can be determined uniquely as functions of $c_{(I)}$, $c_{(K)}$.
551:
552: For $\beta\sqrt{-\Lambda}|v|>k$, the functions $\BesI$ grow exponentially,
553: whereas the functions
554: $\BesK$ exponentially decay. In consequence, only the eigenfunctions
555: with $c_{(I)}=0$ will contribute to the spectral decomposition of
556: $\ub{\Theta}$. This implies that the spectrum of $\ub{\Theta}$ equals
557: $\Sp(\ub{\Theta})=[0,\infty)$ and is continuous. Furthermore, due to
558: \eqref{eq:WDWelimit}, the eigenfunctions with $c_{(I)}=0$ are Dirac
559: delta normalizable. Therefore, we can choose the basis setting
560: $\ub{e}_{\omega}:=\alpha(\omega)\BesK_{ik}(\beta\sqrt{-\Lambda}|v|)$,
561: where $\alpha$ is a real, positive, $\omega$-dependent normalization
562: factor chosen to satisfy the relation
563: \begin{equation}
564: \sip{\ub{e}_{\omega}}{\ub{e}_{\omega'}}\ =\ \delta(\omega,\omega') \ .
565: \end{equation}
566:
567: % positive and negative frequency
568: At this point, we note that the structure of the spectral decomposition of
569: $\ub{\Theta}$ is similar to the one of the WDW limit for the $k=1$ FRW model
570: \cite{apsv}, so that we can follow the construction used there. Each
571: element $\psi(v)$ of $L^2(\re,\ub{B}(v)\rd v)$ can be decomposed in the
572: basis $\ub{e}_{\omega}$:
573: \begin{equation}
574: \psi(v)\ =\ \int_0^{\infty} \rd \omega \tilde{\psi}(\omega)
575: \ub{e}_{\omega}(v) \ .
576: \end{equation}
577: where $\tilde{\psi}\in L^2(\re,\rd\omega)$. Therefore, the solutions
578: to the evolution equation \eqref{eq:WDWmain} with initial data
579: in the Schwartz space can be represented in terms of the two functions
580: $\Psi_{\pm}(\omega)\in L^2(\re,\rd\omega)$
581: \begin{equation}\label{eq:WDWgensol}
582: \Psi(v,\phi)\ =\ \int\rd\omega\left[
583: \tilde{\Psi}_+(\omega)\ub{e}_{\omega}(v)e^{i\omega\phi} +
584: \tilde{\Psi}_-(\omega)\bar{\ub{e}}_{\omega}(v)e^{-i\omega\phi}
585: \right] \ .
586: \end{equation}
587: The solutions with vanishing $\ub{\tilde{\Psi}}_{+}$ and
588: $\ub{\tilde{\Psi}}_{-}$ (denoted in the following as $\ub{\Psi}_-$,
589: $\ub{\Psi}_+$) are called the negative and positive frequency solutions
590: respectively. Their general form can be written in terms of the square
591: root of the $\ub{\Theta}$ operator; namely, for initial data $\psi_o(v)$
592: specified at $\phi=\phi_o$, we have:
593: \begin{equation}\label{eq:WDWmainpm}
594: \Psi_{\pm}(v,\phi)\ =\ e^{\pm i\sqrt{\ub{\Theta}}(\phi-\phi_o)}\psi_o(v) \ .
595: \end{equation}
596:
597:
598: \subsection{Physical Hilbert space, observables}
599: \label{sec:WDWphys}
600:
601: % construction of Hilbert space
602: To construct the physical Hilbert space $\ubHilp$, we again follow
603: \cite{aps-imp,apsv}. As $\ub{\Theta}$ is the sum of the $\ub{\Theta}_o$
604: operator (which is just $\partial_{\ln|v|}^2$) and the positive potential
605: term, it is essentially self-adjoint and positive-definite \cite{klp-nL}.
606: Friedrich's extension of it is thus a unique self-adjoint one. One can
607: then apply group averaging techniques \cite{gave} (see the discussion in
608: \cite{aps-old}) to find $\ubHilp$ and the inner product. The result is the
609: following: the space $\ubHilp$ itself consists of normalizable
610: solutions to \eqref{eq:WDWmain}; however, as the spaces of positive
611: and negative frequency solutions are superselected sectors, we can
612: take as $\ubHilp$ the restriction to just one of them. Following
613: previous works, we chose positive frequency part, thus defining
614: $\ubHilp$ as:
615: \begin{equation}
616: \ub{\Psi}(v,\phi)\ =\ \int\rd\omega\tilde{\ub{\Psi}}(\omega)
617: \ub{e}_{\omega} e^{i\omega\phi} \ , \qquad \
618: \tilde{\ub{\Psi}}\in L^2(\re^+,\rd\omega) \ .
619: \end{equation}
620: The physical inner product within this space can be written as:
621: \begin{equation}
622: \sip{\ub{\Psi}}{\ub{\Phi}}\
623: =\ \int_{\phi=\phi_o} \ub{B}(v)\rd v \bar{\ub{\Psi}}(v)\ub{\Phi}(v) \ .
624: \end{equation}
625:
626: %observables
627: In order to be able to extract physical information out of our system,
628: we need to define a set of Dirac observables, i.e. self-adjoint operators
629: preserving $\ubHilp$. Here again we can directly use the scalar field
630: momentum $\hat{p}_{\phi}$ and $|\hat{v}|_{\phi}$ , the amplitude of $v$ at
631: a given $\phi$, defined already for $\Lambda=0$ and $k=1$. Their
632: action on the elements $\Psi$ of $\ubHilp$ is the following
633: \begin{subequations}\label{eq:WDWobs}\begin{align}
634: \hat{p}_{\phi}\Psi\ &=\ -i\hbar\partial_{\phi}\Psi \ , &
635: |\hat{v}|_{\phi'}\Psi\
636: &=\ e^{i\sqrt{\ub{\Theta}}(\phi-\phi')}|v|\Psi(v,\phi') \ ,
637: \tag{\ref{eq:WDWobs}}
638: \end{align}\end{subequations}
639: and their expectation values equal respectively:
640: \begin{subequations}\label{eq:WDWexpect}\begin{align}
641: \sbra{\ub{\Psi}}\hat{p}_{\phi}\sket{\ub{\Psi}}\ &= \
642: -i\hbar\int_{\phi=\const}\ub{B}(v)\rd v \bar{\ub{\Psi}}(v,\phi)
643: (\partial_{\phi}\ub{\Psi})(v,\phi) \ , \\
644: \sbra{\ub{\Psi}}|\hat{v}|_{\phi}\sket{\ub{\Psi}}\ &= \
645: \int\ub{B}(v)\rd v |v||\ub{\Psi}(v,\phi)|^2 \ .
646: \end{align}\end{subequations}
647:
648:
649: \subsection{Semiclassical states}
650: \label{sec:WDWstates}
651:
652: Once we have the physical Hilbert space, the inner product and the
653: observables, we
654: can investigate the evolution of a universe represented by a given
655: state. A particularly interesting question one can ask is
656: whether, in the considered system, the singularity is resolved. To address
657: this question, we construct a Gaussian state which, at a given time
658: $\phi_o$, is sharply peaked at a large scalar field momentum
659: $p_{\phi}^{\star}=\hbar\omega^{\star}$ (with spread $\sigma/\sqrt{2}$)
660: and volume $v^{\star}$ and is expanding:
661: \begin{equation}\label{eq:WDWpsi}
662: \Psi(v,\phi)\ =\ \int_0^{\infty}\rd\omega \,
663: e^{-\frac{(\omega-\omega^{\star})^2}{2\sigma^2}} \,
664: \ub{e}_{\omega}(v) \,
665: e^{i\omega(\phi-\phi^{\star})} \ ,
666: \end{equation}
667: where
668: \begin{equation}\label{eq:phistar}
669: \phi^{\star}= \frac{1}{\sqrt{12\pi G}}\arcosh\left(
670: \frac{3K\sqrt{12\pi G}}%
671: {(4\pi\gamma\lPl^2)^{3/2}} \frac{p_{\phi}^{\star}}%
672: {\sqrt{|\Lambda|}v^{\star}} \right) + \phi_o \ .
673: \end{equation}
674:
675: Because of the complicated form of $\ub{e}_{\omega}$, the wave function
676: \eqref{eq:WDWpsi} and expectation values \eqref{eq:WDWexpect} were
677: calculated numerically (see section \ref{sec:num} for the details). An
678: example of the results is shown on Figs \ref{fig:WDW}. The state remains
679: semiclassical (sharply peaked) and simply follows the classical trajectory
680: \eqref{eq:pre-class-traj}
681: \begin{equation}\label{eq:WDWtraj}
682: v(\phi)\ =\ \frac{3K\sqrt{12\pi G}}{(4\pi\gamma\lPl^2)^{\frac{3}{2}}}
683: \,\frac{p_{\phi}^{\star}}{\sqrt{|\Lambda|}}\,
684: \left[\cosh\left(\sqrt{12\pi G}(\phi-\phi^{\star}+\phi_o)
685: \right)\right]^{-1} \
686: \end{equation}
687: to the big bang and big crunch singularities. In consequence,
688: similarly to the $\Lambda=0$ case, the classical singularities are not
689: resolved.
690: \begin{figure}[tbh!]\begin{center}
691: $a)$\hspace{8cm}$b)$
692: \includegraphics[width=3.2in,angle=0]{plots/WDW-3d.eps}
693: \includegraphics[width=3.2in,angle=0]{plots/WDW-traj.eps}
694: \caption{An example of a Wheeler-DeWitt Gaussian wave packet
695: \eqref{eq:WDWpsi} generated for the parameter values
696: $\Lambda=-0.01$, $p_{\phi}^{\star}=5\cdot 10^3$, $\Delta
697: p_{\phi}/p_{\phi}^{\star}=0.02$ and $\phi^{\star}=0$. Fig.$a)$ shows the
698: absolute value of the wave function. For the presentation clarity, only
699: the points of $|\Psi(v,\phi)|>10^{-6}$ were plotted. Fig.$b)$
700: presents the expectation values and dispersions of
701: $|\hat{v}|_{\phi}$ (red bars) compared against the classical
702: trajectory $v(\phi)$ (blue line). As we can see, the quantum
703: trajectory agrees with the classical one (the difference being much
704: smaller that the spread). Due to the large changes in magnitude of $v$
705: during the evolution, the trajetory was plotted in logarythmic scale.}
706: \label{fig:WDW}
707: \end{center}\end{figure}
708:
709:
710:
711: \section{Physical sector of LQC}
712: \label{sec:LQC}
713:
714: The analysis in the previous section allowed to find dynamics
715: predicted by the WDW limit of the considered LQC model. Now we perform
716: an analogous study of the model of interest. Due to qualitative
717: similarities of the Hamiltonian constraint with its WDW limit, the
718: analysis can be performed analogously to the one done in section
719: \ref{sec:WDW} (with only slight modifications required by the fact that
720: $\Theta$ is now a difference operator). Following that work, we again
721: restrict the study to states symmetric under parity
722: reflection.\footnote{It is also correct to work with the antisymmetric
723: sector of the theory. We discuss that case in Appendix \ref{sec:antisymm}.}
724:
725: First we note that, thanks to the fact that $\Theta$ is a difference
726: operator, we can naturally divide the gravitational kinematical
727: Hilbert space onto superselected sectors $\Hilkg =
728: \bigoplus_{\varepsilon\in[0,2]} \Hilkge$, where $\Hilkge$ are the
729: restrictions of $\Hilkg$ to the functions supported on the sets
730: $\lat_{\varepsilon} := \{\pm\varepsilon+4n;\ n\in\integ\}$ preserved by the
731: action of the Hamiltonian constraint \eqref{eq:Theta} and parity reflection
732: $\Pi: \psi(v)\mapsto \psi(-v)$. Following the literature, we call these
733: sets {\it lattices} and work with single sectors
734: $\Hilke:=\Hilkge\otimes\Hilkf$. The kinematical inner product
735: corresponding to them is just a restriction of the product of $\Hilk$.
736:
737: For each of the sectors illustrated above, the operator $\Theta$ is
738: obviously well defined and symmetric (with respect to the measure
739: $B(v)\rd\mu_{\Bohr}$) on the domain $D_{\varepsilon}$ -- the space of finite
740: combinations of $\sket{v}$ with $v\in\lat_{\varepsilon}$. Its
741: mathematical properties were rigorously analyzed in \cite{klp-nL}. It
742: is essentially self-adjoint, its extension is positive-definite and
743: its spectrum is discrete.
744: The first two properties allow us again to choose $\phi$ as an emergent
745: time and treat $\Theta$ as an evolution operator.
746:
747: The discreteness of $\Theta$'s spectrum implies that the eigenfunctions
748: relevant for its spectral decomposition are normalizable. Furthermore,
749: a numerical study (discussed in section \ref{sec:num}) shows that
750: the spectrum is nondegenerate. In consequence, for each allowed value
751: of the label $\varepsilon$, we can build the physical Hilbert space
752: $\Hilpe$ as a space of normalizable positive frequency solutions to
753: \eqref{eq:Theta}, analogously to the construction in sections
754: \ref{sec:WDWsols} and \ref{sec:WDWphys}:
755: \begin{equation}\label{eq:Psi}
756: i\partial_{\phi}\Psi\ =\ \sqrt{\Theta}\Psi \ , \qquad
757: \Psi(v,\phi)\ =\ \sum_{n\in\natu} \tilde{\Psi}_n\, e_n(v)\,
758: e^{i\omega_n\phi} \ ,
759: \end{equation}
760: where $\Tilde{\Psi}$ are square summable and $e_n(v)$ are symmetric
761: in $v$ and normalized eigenfunctions of $\Theta$, corresponding to
762: eigenvalues $\omega_n^2$ which form the basis of $\Hilpe$.
763: The physical inner product can be found through group averaging
764: analogously to the WDW case and can be written in the form
765: \begin{equation}\label{eq:IP}
766: \sip{\Psi}{\Phi}\ =\ \sum_{n=0}^{\infty}
767: \bar{\tilde{\Psi}}_n\tilde{\Phi}_n\
768: =\ \sum_{v\in\lat_{\epsilon}} B(v)\bar{\Psi}(v)\Phi(v) \ .
769: \end{equation}
770:
771: % - observables
772: To complete the quantization program we need to choose a set of
773: Dirac observables. In order to be able to compare the results with the WDW
774: limit, we choose the operators analogous to \eqref{eq:WDWobs}
775: \begin{subequations}\label{eq:obs}\begin{align}
776: \hat{p}_{\phi}\Psi\ &=\ -i\hbar\partial_{\phi}\Psi \ , &
777: |\hat{v}|_{\phi'}\Psi\
778: &=\ e^{i\sqrt{\Theta}(\phi-\phi')}|v|\Psi(v,\phi') \ .
779: \tag{\ref{eq:obs}}
780: \end{align}\end{subequations}
781: Their expectation values are equal respectively to
782: \begin{subequations}\label{eq:expect}\begin{align}
783: \sbra{\Psi}\hat{p}_{\phi}\sket{\Psi}\
784: &=\ -i\hbar
785: \sum_{v\in\lat_{\epsilon},\ \phi=\const} B(v)
786: \bar{\Psi}(v,\phi) (\partial_{\phi}\Psi)(v,\phi) \ , \\
787: \sbra{\Psi}|\hat{v}|_{\phi}\sket{\Psi}\
788: &=\
789: \sum_{v\in\lat_{\epsilon}} B(v) |\Psi(v,\phi)|^2 \ .
790: \end{align}\end{subequations}
791:
792: To calculate an explicit form of $\Psi$ (needed to find the
793: expectation values) one needs to find the spectrum of $\Theta$ and the
794: explicit form of its normalizable eigenfunctions. Because of the
795: complicated structure of $\Theta$, in order to do so one needs to resort to
796: numerical methods. We present them in the next section.
797:
798:
799:
800: \section{Numerical study}\label{sec:num}
801:
802: This section is divided onto two parts. In section \ref{sec:num-eig},
803: we present the methods and results of identifying the spectrum of the
804: $\Theta$ operator and finding normalizable eigenfunctions. The
805: techniques for computing the wave function and the expectation values
806: are presented in section \ref{sec:num-state}. In both parts, we
807: applied the (appropriately refined) methods used already for the $k=1$
808: model and introduced in \cite{apsv}. Unless specified otherwise, from
809: now on we will work with units in which $G=1$.
810:
811: \subsection{Spectrum of $\Theta$}\label{sec:num-eig}
812:
813: In order to construct the Hilbert space $\Hilpe$, one needs to find the
814: eigenfunctions supported on the lattice $\lat_{\varepsilon}$, which
815: consists of two sublattices
816: $\lat_{\pm|\varepsilon|}:=\{\pm|\varepsilon|+4n;\ n\in\integ\}$
817: invariant with respect to the action of the Hamiltonian constraint. Each of
818: such eigenfunctions (denoted here as $\psi$) is a solution to a
819: difference equation:
820: \begin{equation}\label{eq:eigen}
821: -\omega^2 B(v)\psi(v)\ =\ C^+(v)\psi(v+4)
822: + (C^o(v)+C_{\Lambda}(v))\psi(v) + C^-(v)\psi(v-4) \ ,
823: \end{equation}
824: where $\omega^2$ is the eigenvalue that each given eigenfunction
825: corresponds to and $C^o, C^{\pm}, C_{\Lambda}$ are given by
826: (\ref{eq:BCLcoeff},~\ref{eq:ThCoeffdef}). On each sublattice, this is a
827: second-order equation --
828: one needs to specify the initial data at two neighboring points ($v_{\init}$,
829: $v_{\init}+4$) to uniquely define a solution. The symmetry condition
830: $\psi(v)=\psi(-v)$, however, restricts the amount of initial data in
831: the following way:
832: \begin{enumerate}[(i)]
833: \item For $\varepsilon\in (0,2)$, the sublattices $\lat_{\pm|\varepsilon|}$
834: are disjoint and the parity reflection $\Pi$ transforms one onto another.
835: Therefore one needs to specify an initial data $\psi(v_{\init})$,
836: $\psi(v_{\init}+4)$ for just one of them, say
837: $\lat_{+|\varepsilon|}$, and complete it by the action of $\Pi$. We denote
838: such lattices as {\it generic}.
839: \item When $\varepsilon=0,2$ the sublattice $\lat_{+|\epsilon|}$
840: coincides with $\lat_{-|\epsilon|}$ and is invariant with respect
841: to parity reflection. The condition $\psi(v)=\psi(-v)$, applied to
842: \eqref{eq:eigen}, imposes on it an additional constraint of the
843: form depending on the value of $\varepsilon$:
844: \begin{itemize}
845: \item $\varepsilon=0$: $\psi(-4)=\psi(0)=\psi(4)$,
846: \item $\varepsilon=2$: $\psi(-2)=\psi(2)$. Here the equality
847: $C^-(2)=C^+(-2)=0$ implies additionally $\psi(\pm 6) =
848: -[(\omega^2B(2)+C_{\Lambda}(2)+C^o(2))/C^+(2)]\psi(\pm 2)$.
849: \end{itemize} In consequence, the value of $\psi$ at just one point
850: ($v=0$ or $v=2$) determines the entire eigenfunction. These cases are
851: called {\it exceptional}.
852: \end{enumerate}
853:
854: % preliminary results
855: The degrees of freedom specified above are complex; however, since
856: the coefficients of \eqref{eq:eigen} are real, $\psi$ satisfies it
857: iff so do its components $\Re(\psi)$, $\Im(\psi)$. Therefore, we can
858: safely restrict our study to a real $\psi$.
859:
860: Upon this restriction, the space of solutions to \eqref{eq:eigen}
861: is $1$-dimensional for exceptional lattices and $2$-dimensional for
862: generic ones. Once the initial data are specified appropriately for
863: each case, the function $\psi$ can be found by solving \eqref{eq:eigen}
864: iteratively.
865:
866: To determine the properties of $\psi$, we calculated the solutions in
867: a wide range of both $\Lambda$ ($[-10,-10^{-6}]$) and $\omega$
868: ($[0,10^5\hbar]$). The qualitative features of the found solutions is
869: visualized on fig.\ref{fig:eig-exampl}; in general, for each $\psi$ one
870: can distinguish $5$ zones of distinct behavior, and the boundaries of
871: these zones are specified by the functions $v_B(\omega)$ and $v_R(\omega)$,
872: approximately equal to, respectively, the position of the bounce for a
873: $\Lambda=0$ universe with $p_{\phi}^{\star} = \hbar\omega$ (determined in
874: \cite{aps-imp}) and the value of $v$ at the recollapse point of the classical
875: universe (given by \eqref{eq:WDWtraj} at $\phi=\phi^{\star}-\phi_o$).
876: \begin{enumerate}[(i)]
877: \item For $|v|<v_B(\omega)$, the amplitude of $\psi$ grows/decays
878: quasi-exponentially. \label{it:z1}
879: \item For $v_B<|v|<v_R$, the behavior of $\psi$ is oscillatory
880: (similar in nature to the behavior of \eqref{eq:WDWgeneig}). \label{it:z2}
881: \item When $|v|>v_R$ the eigenfunction grows/decays exponentially
882: with $|v|$ (where the exponential growth is a generic behavior).
883: \label{it:z3}
884: \end{enumerate}
885: Note that for small $\omega$, the zones \eqref{it:z1} and
886: \eqref{it:z2} may be empty (see fig.\ref{fig:eig-low} for examples).
887:
888: \begin{figure}[tbh!]\begin{center}
889: $a)$\hspace{8cm}$b)$
890: \includegraphics[width=3.2in,angle=0]{plots/eig-exampl-0.eps}
891: \includegraphics[width=3.2in,angle=0]{plots/eig-exampl-1.eps}
892: \caption{Examples of eigenfunctions of $\Theta$ supported on the
893: lattices $\lat_{\varepsilon}$ for $\varepsilon=0$ $(a)$ and $\varepsilon=1$
894: $(b)$.\\
895: $a)$ shows a normalizable eigenfunction of $\omega\approx300.45$
896: (red) and two divergent ones of $\omega$ respectively smaller
897: (green) and larger (blue) by $0.1$. For clarity, only the positive $v$
898: part is shown.\\
899: $b)$ presents the absolute value of a normalizable eigenfunction of
900: $\omega\approx52.85$ (red) along with two divergent examples: generic
901: (green) and left-converging (blue) generated for, respectively
902: $\omega\approx53.35$ and $54.35$. To show the behavior in a wide range
903: of values, a logarythmic scale was used for the $y$-axis.\\
904: Both figures correspond to $\Lambda=-0.01$.}
905: \label{fig:eig-exampl}
906: \end{center}\end{figure}
907:
908: Since we search for normalizable functions only, we have to select the
909: ones which decay exponentially in the zones of type \eqref{it:z3}. We
910: identify them numerically using different methods depending on whether
911: the eigenfunctions are supported on generic or exceptional lattices.
912:
913: % exceptional case
914: On exceptional lattices, each eigenfunction $\psi_{\omega}$ is (for
915: a given $\omega$) determined uniquely up to a global scaling. To find the
916: normalizable solutions, we scan the domain of $\omega$ using the
917: following observation:
918:
919: \begin{obs}
920: For a chosen $\omega\in [\omega_{1},\omega_{2}]$,
921: $\psi_{\omega}(\epsilon)=1$ and $v\gg v_R(\omega_2)$, the value
922: $\psi_{\omega}(v)$ is a continuous function of $\omega$ (more specifically,
923: a polynomial) and its sign changes
924: quasi-periodically. Furthermore, if we define $\omega_{v,n}$ as the
925: values of $\omega$ such that $\Psi_{\omega_{v,n}}(v) = 0$, the limits
926: $\omega_n := \lim_{v\to\infty}\omega_{v,n}$ are well defined and
927: correspond to the values of $\omega$ for which $\psi_{\omega}$
928: decays in zone \eqref{it:z3}.
929: \end{obs}
930:
931: In practice, due to the precision bound posed by numerical round-off,
932: it is enough to (instead of finding the limits) look for values
933: of $\omega_{v,n}$ at $v_T$ sufficiently far away from $v_R$. For the
934: actual search, we selected $v_T=\max(2000,1.3v_R)$. The search itself
935: was performed in two steps:
936: \begin{itemize}
937: \item First the sign of $\Psi_{\omega}(v_T)$ was checked for
938: values of $\omega$ uniformly separated by a distance around $0.1$.
939: \item If a change of sign was detected between neighboring points, the
940: value of $\omega_{n,v_T}$ was found via bisection.
941: \end{itemize}
942:
943: % generic case
944: For generic lattices, the space of solutions is, up a to global rescaling,
945: $1$-dimensional, so besides $\omega$ we need to specify the value of $\psi$
946: at two points $v_I,v_I+4\in\lat_{+|\epsilon|}$. An additional
947: complication is the fact that now the behavior in zones of type \eqref{it:z3}
948: for $v>0$ and $v<0$ is independent. The function may grow for positive $v$
949: while decaying for negative ones and vice versa. Therefore, to find the
950: desired functions we divide the search procedure onto two steps:
951: \begin{itemize}
952: \item First we identify the family $\psi_{\omega}$ of functions
953: decaying in zone \eqref{it:z3} for $v<0$ (further denoted as {\it
954: left-converging}). To do so, we parametrize the initial data at
955: $v_I,v_I+4$ by a parameter $\alpha\in[0,\pi]$
956: \begin{subequations}\label{eq:ida}\begin{align}
957: \psi_{\alpha,\omega}(v_I)\ &=\ \cos(\alpha) \ , &
958: \psi_{\alpha,\omega}(v_I+4)\ &=\ \sin(\alpha) \ ,
959: \tag{\ref{eq:ida}}\end{align}\end{subequations}
960: and scan the domain of $\alpha$ for the values at which the limit
961: $\lim_{v\to-\infty}\psi_{\alpha,\omega}(v) = 0$. Analogously to
962: the exceptional lattice case, it is enough here to just choose some
963: value $-v_R(\omega)\gg v_{T^-}\in\lat_{+|\varepsilon|}$ and look
964: for the values of $\alpha$ at which $\psi_{\alpha,\omega}(v_{T^-})=0$.
965: In practice, it suffices to choose $v_{T^-}\approx -v_T$, where
966: $v_T$ is the value defined for exceptional lattices. The scan
967: method is analogous to the scan of $\omega$ in the exceptional case: we
968: divide the domain of $\alpha$ into $10$ uniform intervals and if
969: a change of sign of $\psi_{\alpha,\omega}(v_{T^-})$ is detected
970: within an interval, the precise value of $\alpha$ is found via
971: bisection.
972:
973: It was checked by inspection that, for each $\omega$, there is
974: exactly one value of $\alpha$ satisfying the above requirement. In
975: consequence, for each $\omega$ the eigenspace of left-converging
976: functions is $1$-dimensional.
977: \item Once the family $\psi_{\omega}$ of left-converging functions
978: is selected, we choose some $v_T\approx
979: v_{T^+}\in\lat_{+|\varepsilon|}$ and scan the domain of $\omega$
980: for values at which $\psi_{\omega}(v_{T^+})=0$, via the method
981: specified for exceptional lattices.
982: \end{itemize}
983:
984: % result for small w -- application of initial data
985: \begin{figure}[tbh!]\begin{center}
986: $a)$\hspace{8cm}$b)$
987: \includegraphics[width=3.2in,angle=0]{plots/eigenfunct-0.eps}
988: \includegraphics[width=3.2in,angle=0]{plots/eigenfunct-1.eps}
989: \caption{The eigenfunctions $e_0$ to $e_4$ of $\Theta$, corresponding to
990: $\Lambda=-0.01$ and supported on $\lat_{\varepsilon}$ with
991: $\varepsilon=0$ $(a)$ and $\varepsilon=1$ $(b)$. For clarity, only
992: the $v>0$ part was shown in $a)$ and only the part supported on
993: $\lat_{+|\varepsilon|}$ was shown in $b)$.}
994: \label{fig:eig-low}
995: \end{center}\end{figure}
996:
997: The search was first performed for small $\omega$ ($<50$) to find the
998: qualitative behavior of normalizable eigenfunctions. An example of the
999: results is shown in Figs.~\ref{fig:eig-low} and \ref{fig:dw}a. All
1000: found eigenfunctions belong to one of the following groups:
1001: \begin{enumerate}[(1)]
1002: \item Suppressed on the $v<0$ side with suppression exponential in
1003: $\omega$.
1004: \item Suppressed for $v>0$.
1005: \item Peaked about $v=0$.
1006: \end{enumerate}
1007: In consequence, it is most convenient, from the point of view of
1008: the numerical precision of the solutions, to specify the initial data at
1009: $v_I\approx\pm v_R$. However, because of the quasi-exponential behavior of
1010: the eigenfunctions in zone \eqref{it:z1}, we can calculate (with a sufficiently
1011: small numerical error) only the solutions suppressed on
1012: the side where the initial data were specified. Therefore it is
1013: necessary to repeat the search twice: for $v_I\approx v_R$ and
1014: $v_I\approx -v_R$.
1015:
1016: \begin{figure}[tbh!]\begin{center}
1017: $a)$\hspace{8cm}$b)$
1018: \includegraphics[width=3.2in,angle=0]{plots/Eigenvals.eps}
1019: \includegraphics[width=3.2in,angle=0]{plots/DwLimit.eps}
1020: \caption{$a)$ The lowest ($\omega<44$) elements of $\Theta$'s spectrum are
1021: shown as functions of $\pm|\varepsilon|$. The eigenvalues are divided
1022: into three groups corresponding to the following eigenfuctions:
1023: $(1)$ left-suppressed (red crosses), for which $>50\%$ of
1024: the norm is located on $v>0$, $(2)$ right-suppressed (green x-es)
1025: defined analogously and $(3)$ singularity-peaked (blue stars),
1026: where $>50\%$ of the norm is located at the three points closest
1027: to $v=0$. \\
1028: $b)$ The large $\omega$ limit of the eigenvalue separation
1029: $\Delta\omega$, shown as function of $\Lambda$ (red crosses). The
1030: blue line represents the small $\Lambda$ limit given by \eqref{eq:dwlow}.}
1031: \label{fig:dw}
1032: \end{center}\end{figure}
1033:
1034: % THE RESULTS of spectrum search
1035: The spectrum scan described above was performed for $18$
1036: values of $\Lambda$ ranging from $-20$ to $-10^{-6}$. It revealed
1037: the following properties (visualized in figs.~\ref{fig:eig-low} --
1038: \ref{fig:A}).
1039: \begin{itemize}
1040: \item As analytically predicted, for each $\epsilon$ the spectrum of
1041: $\Theta$ is discrete and the eigenvalues are
1042: isolated. With the exception of the lowest $\omega$, the eigenfunctions are
1043: highly (exponentially in $\sqrt{\omega}$) suppressed for one triad
1044: orientation (sign of $v$). The eigenvalues corresponding to them
1045: are continuous functions of $\epsilon$. The density of eigenvalues
1046: is twice higher on the generic lattices than on the exceptional
1047: ones. Furthermore, for $\varepsilon=2$, the two families of left-suppressed
1048: and right-suppressed eigenfunctions converge (see fig.~\ref{fig:dw}a).
1049: \item The separation $\Delta\omega_{n}:=\omega_{n+1}-\omega_n$ is
1050: not uniform. It depends on $\epsilon$ and $\Lambda$ as well as
1051: $n$. However, for large values of $\omega$, $\Delta\omega_n$
1052: converges to the limit value $\Delta\omega$ with convergence rate
1053: $\omega^{-2}$ (see fig.\ref{fig:A}a)
1054: \begin{equation}\label{eq:dw}
1055: \Delta\omega_n\ =\ \Delta\omega + O(\omega^{-2}) \ , \qquad
1056: \Delta\omega\ =\ \lim_{n\to\infty}\Delta\omega_n \ .
1057: \end{equation}
1058: Numerical inspection shows that the correction satisfies (with
1059: the exception of the lowest $\omega$) the following bound relation
1060: \begin{equation}\label{eq:A-bound}
1061: |\Delta\omega-\Delta\omega_n|\ \leq\
1062: \frac{A (\Delta\omega)^2}{\omega^{2}} \ ,
1063: \end{equation}
1064: where, for $|\Lambda|<10$, $A<0.21$ and $A$ decreases for smaller
1065: $|\Lambda|$, reaching in the $|\Lambda|\to 0$ limit the value $A
1066: \approx 0.1358\pm 2\cdot 10^{-4}$ (see fig.\ref{fig:A}).
1067: \item The limit $\Delta\omega$ was found numerically via $4$th order
1068: polynomial extrapolation of $\Delta\omega_n$.
1069: It is a function of $\Lambda$
1070: only, i.e. it does not depend on $\epsilon$. Its values
1071: for different superselection sectors agree up to $10^{-9}$
1072: precision. The dependence on $\Lambda$ found numerically is shown
1073: on fig.~\ref{fig:dw}b. For small values of $|\Lambda|$ it can be
1074: approximated via a power function
1075: \begin{equation}\label{eq:dwlow}
1076: \Delta\omega \approx a |\Lambda G|^b
1077: \end{equation}
1078: where $a\approx3.87$ and $b\approx0.0489$.
1079: \end{itemize}
1080:
1081: \begin{figure}[tbh!]\begin{center}
1082: $a)$\hspace{8cm}$b)$
1083: \includegraphics[width=3.2in,angle=0]{plots/A-range.eps}
1084: \includegraphics[width=3.2in,angle=0]{plots/A-limit.eps}
1085: \caption{The rescaled eigenvalue separation correction term
1086: $A(\omega):=|\Delta\omega_n-\Delta\omega|
1087: (\omega_n/\Delta\omega)^2$ is shown in $a)$ for several values of
1088: $\Lambda$. Its $\omega\to\infty$ limit is plotted in $b)$ as a
1089: function of $\Lambda$. The 'wiggles' at small values of
1090: $|\Lambda|$ are the results of numerical errors due to the precision
1091: limitation of the applied calculation method.}
1092: \label{fig:A}
1093: \end{center}\end{figure}
1094:
1095:
1096: The spectrum and normalizable eigenfunctions found here may be next
1097: used to construct the semiclassical states. Details of this construction
1098: will be presented in the next section.
1099:
1100:
1101: \subsection{Semiclassical states, evolution}\label{sec:num-state}
1102:
1103: Once we know the values of $\omega_n$ and $e_n(v)$, the construction of
1104: a physical state from \eqref{eq:Psi} is straightforward. There are two
1105: possibilities here: direct summation of equation
1106: (\ref{eq:Psi}b) or numerical integration via equation
1107: (\ref{eq:Psi}a) (or equivalently via \eqref{eq:Theta}) of some initial data
1108: specified at a given $\phi_o$. To find these data, we again have two
1109: methods at our disposal: one of them is the same direct summation of
1110: (\ref{eq:Psi}b), but applied to one slice, whereas the second
1111: possibility is the use of a slice of a WDW semiclassical state (see
1112: section \ref{sec:WDW}) peaked at
1113: large $v^{\star}$, where we do not expect strong quantum-geometric
1114: effects. In practice we used the second method, integrating the state
1115: in $\phi$ via equation \eqref{eq:Theta} and using as initial data both
1116: the WDW slices and the results of the summation of (\ref{eq:Psi}b). The first
1117: method of state calculation was used only to measure the wave packet
1118: spread increase in large intervals of $\phi$, as the integration
1119: methods were not precise enough for this application.
1120:
1121: \subsubsection{Initial data}
1122:
1123: Let us focus on the second method of initial data specification: constructing
1124: the WDW slice.
1125: In order to be able to directly compare the dynamics of LQC model and
1126: its WDW limit described in section \ref{sec:WDW} we take as the
1127: initial data the $\phi=\phi_o$ section of the Gaussian state
1128: \eqref{eq:WDWpsi} peaked at $p^{\star}_{\phi}=\hbar\omega^{\star}$ and
1129: $v^{\star}$. Since \eqref{eq:Theta} is a second order equation, to
1130: specify the initial data completely we also need $\dot{\Psi}(v,\phi_o)$
1131: -- the first order derivative of $\Psi$ with respect to $\phi$. We get
1132: it by integrating the integrand of \eqref{eq:WDWpsi},
1133: multiplied by $i\omega$, over $\omega$.
1134:
1135: In order to calculate the specified integrals, we first need to compute
1136: the values of $\ub{e}_{\omega}(v)$, which are the normalized Bessel functions
1137: $\BesK$ (see section \ref{sec:WDWsols}). To do so, we apply the
1138: combined method specified by Gil, Segura and Temme \cite{BesK}.
1139:
1140: Once we have $\ub{e}_{\omega}(v)$, we integrate \eqref{eq:WDWpsi} (and
1141: the analogous expression for $\dot{\Psi}$) over the domain
1142: $[\omega^{\star}-7\sigma, \omega^{\star}+7\sigma]$, using the trapezoid
1143: method. Such choice of domain provides sufficient precision -- the errors
1144: due to the removed tails are much smaller than the error associated
1145: with the computation of the $\BesK$ functions.
1146:
1147: % note on negative frequency correction
1148: Note that we intend to construct the initial data corresponding to the
1149: positive frequency solution to \eqref{eq:Theta}. In that case,
1150: $\dot{\Psi}$ is already determined by $\Psi$ via (\ref{eq:Psi}a).
1151: On the other hand, we determined it using positive frequency WDW equation
1152: \eqref{eq:WDWmainpm}. Since $\sqrt{\Theta}$ differs from
1153: $\sqrt{\ub{\Theta}}$, our initial data is not a pure positive frequency
1154: solution. To minimize the negative frequency part, we choose the following
1155: method to construct states sharply peaked at large $v^{\star}$:
1156: we require $v^{\star}$ be greater than $2.5p_{\phi}^{\star}/\hbar$,
1157: which keeps the negative frequency part below $10^{-3}$ of the
1158: entire wave packet norm.
1159:
1160: % direct LQC construction
1161: We avoid the above problem if we use directly the basis of functions
1162: $e_n(v)$ and sum them using (\ref{eq:Psi}b). In that case, as
1163: the spectral profile $\tilde{\Psi}_n$ we choose the restriction of the
1164: Gaussian to $\{\omega_n\}$, that is
1165: \begin{equation}\label{eq:tPsi}
1166: \tilde{\Psi}_n\ =\ e^{-\frac{(\omega_n-\omega^{\star})^2}{2\sigma^2}}
1167: e^{-i\omega_n\phi^{\star}}
1168: \end{equation}
1169: where $\hbar\omega^{\star}=p_{\phi}^{\star}$ is again the location of
1170: the peak in the momentum and $\hbar\sigma/\sqrt{2}$ is its spread.
1171: The parameter $\phi^{\star}$ is determined by the position
1172: $v^{\star}$ of the peak in $v$ and value of $\phi_o$ via \eqref{eq:phistar}.
1173: Similarly to the WDW initial data, we sum only over $\omega_n\in
1174: [\omega^{\star}-7\sigma, \omega^{\star}+7\sigma]$. The derivative
1175: $\dot{\Psi}$ is calculated by summing over the individual terms,
1176: multiplied by $i\omega_n$.
1177:
1178: \subsubsection{Evolution}
1179:
1180: Given some initial data, one can integrate it over some interval
1181: $[\phi_o,\phi_1]$ using equation \eqref{eq:Theta}, which is a system of
1182: a countable number of coupled ordinary differential equations (ODE). Due to
1183: the $v$-reflection symmetry, it is enough to restrict the domain of
1184: integration to $\lat^+=\lat_{+|v|}$ for generic lattices and
1185: $\lat^+=\lat_{\epsilon}\cap\re^{+}$ for exceptional ones.
1186: Additional, the numerical nature of our study requires that we further
1187: restrict the domain of $v$ to the finite subset $\lat^+_{v_{\max}} :=
1188: \lat^+\cap[-v_{\max},v_{\max}]$, imposing at the outermost points of
1189: the domain some (artificial) boundary conditions. Since the system
1190: under consideration is a classically recollapsing one, it is enough to
1191: choose the reflective conditions $\Psi = \dot{\Psi} = 0$. To prevent
1192: their interference with the dynamics, we have chosen $v_{\max}$ to be
1193: not smaller than $1.3v_R(\omega^\star)+2000$.
1194:
1195: Upon the above restriction of the $v$ domain, the equation
1196: \eqref{eq:Theta} becomes a finite system of ODEs. We integrate it
1197: using a $4th$-order adaptive Runge-Kutta method (RK4). To adapt the
1198: steps of integration, we compare solutions corresponding to step
1199: $\Delta\phi$ and $\Delta\phi/2$ and require the difference between
1200: them (at a single $\Delta\phi$ step / two $\Delta\phi/2$ steps) to
1201: satisfy the inequality:
1202: \begin{equation}
1203: \| \Psi_{\Delta\phi} - \Psi_{\Delta\phi/2} \|\ \leq\
1204: \frac{\epsilon\Delta\phi}{|\phi_1-\phi_o|} \|\Psi_{\Delta\phi/2}\|
1205: \ ,
1206: \end{equation}
1207: where $\epsilon$ is a preset global bound. The two solutions are compared
1208: via the following norm
1209: \begin{equation}\label{eq:norm1}
1210: \| \Psi \|\ :=\ \sup_{v\in\lat^+_{v_{\max}}} |\Psi(v,\phi)| \ .
1211: \end{equation}
1212: Since only $|\Psi|$ enters the formulae for the expectation values of
1213: $|\hat{v}|_{\phi}$ and $v^2_{\phi}$, it is also convenient to introduce
1214: an auxiliary metric measuring the error in absolute values only
1215: \begin{equation}\label{eq:norm2}
1216: \| \Psi_1 - \Psi_2 \|_{A}\
1217: =\ \sup_{v\in\lat^+_{v_{\max}}} \big||\Psi_1(v,\phi)|-|\Psi_2(v,\phi)|\big|
1218: \end{equation}
1219: An example of convergence test done with respect to both the norm
1220: \eqref{eq:norm1} and the metric \eqref{eq:norm2} is shown in
1221: fig.\ref{fig:conv}, where the results of integration with different
1222: error bounds $\epsilon$ were compared against the result of polynomial
1223: extrapolation at $\epsilon=0$. As we can see, the integration error is
1224: for $|\Psi|$ at least one order of magnitude smaller than that for
1225: $\Psi$ itself.
1226:
1227: \begin{figure}[tbh!]\begin{center}
1228: \includegraphics[width=3.2in,angle=0]{plots/conv-test.eps}
1229: \caption{Convergence test for the integration method of a
1230: Gaussian wave packet generated with $\Lambda=-0.1$ and peaked at
1231: $p^{\star}_{\phi}=10^3$, with relative $p_{\phi}$ spread $0.05$ and
1232: $v^{\star}=0.5\,v_R(p^{\star}_{\phi})$. The initial data were
1233: specified at $\phi=0$ and evolved till $\phi=1$. The upper (red)
1234: curve shows the norm of the difference $\|\Psi_{(N)}-\Psi\|$ between
1235: the slice $\phi=1$ of the solution $\Psi_{(N)}$ corresponding to
1236: the integration with $N$ steps and the same slice of its $N\to\infty$
1237: limit $\Psi$ (found via $8$th order polynomial extrapolation). The lower
1238: (green) curve shows the analogous difference taken with respect to
1239: the metric \eqref{eq:norm2}.}
1240: \label{fig:conv}
1241: \end{center}\end{figure}
1242:
1243: \subsubsection{Observables}\label{sec:lqc-obs}
1244:
1245: Knowing an explicit form of $\Psi$ at
1246: $\lat^+_{v_{\max}}\times[\phi_o,\phi_1]$, we can complete it to
1247: $(\lat^+_{v_{\max}}\cup\lat^-_{v_{\max}})\times[\phi_o,\phi_1]$ (where
1248: $\lat^-_{v_{\max}}:=\{-v:v\in\lat^+_{v_{\max}}\}$) via reflection and
1249: find the expectation values of the observables \eqref{eq:obs}, restricting
1250: the sums (\ref{eq:IP},~\ref{eq:expect}) to a finite domain
1251: $\lat^+_{v_{\max}}\cup\lat^-_{v_{\max}}$. Their dispersions can be in
1252: turn calculated in the standard way
1253: \begin{subequations}\label{eq:disp}\begin{align}
1254: \expect{\Delta|\hat{v}|_{\phi}}^2\ &=\ \expect{\hat{v}^2_{\phi}}
1255: - \expect{|\hat{v}|_{\phi}}^2 &
1256: \expect{\Delta\hat{\phi}_{\phi}}^2\ &=\ \expect{\hat{\phi}^2_{\phi}}
1257: - \expect{\hat{\phi}_{\phi}}^2 \ .
1258: \tag{\ref{eq:disp}}\end{align}\end{subequations}
1259: where $\expect{\hat{v}^2_{\phi}}$, $\expect{\hat{\phi}^2_{\phi}}$ have
1260: a form analogous to \eqref{eq:expect}.
1261:
1262: In addition to $|\hat{v}|_{\phi}$, $\hat{\phi}_{\phi}$, it is useful to
1263: introduce another family of observables: the regularized energy density
1264: at a given moment of $\phi$
1265: \begin{subequations}\label{eq:rho}\begin{align}
1266: \hat{\rho}_{\phi}\
1267: &:=\ \frac{1}{2\lPl^6}\left(\frac{6}{8\pi\gamma}\right)^3
1268: \hat{p}_{\phi}\,\hat{B}_{\phi}\,
1269: \hat{p}_{\phi}\,\hat{B}_{\phi}
1270: \ , & %\dquad
1271: \hat{B}_{\phi}\Psi\
1272: &:=\ e^{i\sqrt{\Theta}(\phi-\phi')}B(v)\Psi(v,\phi') \ .
1273: \tag{\ref{eq:rho}}\end{align}\end{subequations}
1274: We calculate their expectation values via
1275: \begin{equation}
1276: \sbra{\Psi}\hat{\rho}_{\phi}\sket{\Psi}\ =\ -
1277: \frac{K^2}{2\lPl^6}\left(\frac{6}{8\pi\gamma}\right)^3
1278: \sum_{\lat^+{v_{\max}}\cup\lat^-{v_{\max}}} B(v)
1279: |\Phi(v,\phi)|^2 \ ,\quad \Phi\ =\
1280: \partial_{\phi}(\widehat{|v|^{-1}}_{\phi}\Psi) \ ,
1281: \end{equation}
1282: whereas the dispersions can be derived analogously to \eqref{eq:disp}.
1283:
1284: The above methods for calculating the expectation values were applied to
1285: the wave functions calculated earlier through the RK4 method. We analyzed
1286: the states evolved (integrated) from both WDW and exact LQC Gaussian
1287: wave packets corresponding to $17$ values of $\Lambda$ ranging
1288: from $-20$ to $-10^{-6}$, for $5$ different superselection sectors
1289: covering the full range of $\varepsilon$. The peak in momentum
1290: $\phi^{\star}_{\phi}$ covered the values from $5\cdot 10^2$ to $10^4$ ($10$
1291: values), while its relative spread ranged from $0.01$ to $0.1$.
1292:
1293:
1294: \section{Results and discussion}\label{sec:concl}
1295:
1296: \begin{figure}[tbh!]\begin{center}
1297: \includegraphics[width=6.0in,angle=0]{plots/LQC-3dv1.eps}
1298: \caption{The absolute value of the wave function representing
1299: a Gaussian state \eqref{eq:Psi} generated via backward integration
1300: of an initial profile corresponding to $\Lambda=-1$,
1301: $p_{\phi}^{\star}=5\cdot 10^3\hbar$, $\Delta
1302: p_{\phi}/p_{\phi}^{\star}=0.01$, $v^{\star}=0.6\,v_R(p_{\phi}^{\star})$
1303: and evaluated at $\phi_o=0$. For presentation clarity, only values
1304: $>10^{-6}$ were shown on the plot.}
1305: \label{fig:Psi-3d}
1306: \end{center}\end{figure}
1307:
1308: An example of the results of our numerical investigations is presented
1309: in figs. \ref{fig:Psi-3d} -- \ref{fig:rho-zoom}.
1310: The general properties of the considered model are similar to the ones of
1311: the models previously investigated: $\Lambda=0$ and $k=1$, that is:
1312: \begin{itemize}
1313: \item The states remain sharply peaked for long
1314: evolution times. On each superselection sector and large $\omega$, the
1315: spectrum of $\Theta$ quickly approaches uniformity (with approach
1316: rate $\omega^{-2}$). In consequence, a wave packet sharply peaked
1317: at large $p_{\phi}$ should be almost periodic in $\phi$. This
1318: expectation is confirmed by our numerical results, where already for
1319: $p^{\star}_{\phi}$ of the order of few thousands the departures
1320: from periodicity were undetectable within given precision
1321: of integration.
1322: \item For large volumes (small energy densities), the trajectory of the
1323: expectation values $\expect{|v|_{\phi}}$ agrees with the classical
1324: one given by \eqref{eq:WDWtraj}. In particular, the universe
1325: recollapses at the volume predicted by the classical theory even
1326: for large values of $\Lambda$; this was numerically confirmed
1327: for $|\Lambda|$ up to $20$.
1328: \item Once the expectation value of the energy density approaches
1329: the Planck order, we observe the departures from the classical
1330: theory due to quantum-geometric corrections. The corrections act
1331: effectively like an additional repulsive force, which in
1332: particular causes the bounce at the point where the total energy
1333: density $\expect{\hat{\rho}_{\phi}} + \Lambda/(8\pi G)$ approaches
1334: a critical value $\rho_c\approx 0.82\rho_{\Pl}$, identified
1335: already in \cite{aps-imp}.
1336: \item After the bounce, the universe again enters (another) classical
1337: trajectory repeating the cycle of expansion, recollapse and
1338: contraction till the energy density grows again to Planck scale. In
1339: consequence, the evolution is periodic and, similarly to the $k=1$
1340: case, we are dealing with a cyclic model.
1341: \item The wave packet remains sharply peaked even in the region
1342: where the quantum corrections are strong. In consequence, the evolution
1343: can be described by the classical effective dynamics, similarly to the
1344: $\Lambda=0$ case. Indeed, the comparison of the values of
1345: $\expect{|v|_{\phi}}$ with the effective trajectories corresponding to
1346: the holonomy corrections (see Appendix \ref{sec:eff}) has
1347: shown that they agree up to an error well below $\expect{\Delta v}$.
1348: \end{itemize}
1349:
1350: % v-trajectory + rho-tralectory
1351: \begin{figure}[tbh!]\begin{center}
1352: $a)$\hspace{8cm}$b)$
1353: \includegraphics[width=3.2in,angle=0]{plots/LQC-v-traj.eps}
1354: \includegraphics[width=3.2in,angle=0]{plots/LQC-rho.eps}
1355: \caption{The expectation values (red bars) of $|\hat{v}|_{\phi}$ $(a)$ and
1356: $\hat{\rho}_{\phi}$ $(b)$ are compared against
1357: the classical (red lines) and effective (blue lines) trajectories of
1358: $v(\phi)$ and $\rho_{\phi}(v)$ respectively. The data corresponds
1359: to a Gaussian wave packet \eqref{eq:Psi} with $\Lambda=-0.1$,
1360: $p^{\star}_{\phi}=10^4\hbar$, $\Delta p_{\phi}/p^{\star}_{\phi}=0.012$,
1361: $v^{\star}=0.55\,v_R(p^{\star}_{\phi})$
1362: specified at $\phi_o = 0$ and evolved backwards. Because of the large
1363: changes in magnitude of $\expect{\rho_{\phi}}$, a logarythmic
1364: scale was used for the $y$-axis of fig.$b)$.
1365: }
1366: \label{fig:vr-traj}
1367: \end{center}\end{figure}
1368:
1369: % v-zooms: bounce + recollapse
1370: \begin{figure}[tbh!]\begin{center}
1371: $a)$\hspace{8cm}$b)$
1372: \includegraphics[width=3.2in,angle=0]{plots/LQC-v-z1.eps}
1373: \includegraphics[width=3.2in,angle=0]{plots/LQC-v-z2.eps}
1374: \caption{A detailed picture of the comparizon of
1375: $\expect{|v|_{\phi}}$ against the $v(\phi)$ trajectories presented in
1376: fig.~\ref{fig:vr-traj}a is shown near the bounce $(a)$ and recollapse
1377: $(b)$ points respectively.}
1378: \label{fig:v-zoom}
1379: \end{center}\end{figure}
1380:
1381: % rho-zooms: bounce + recollapse
1382: \begin{figure}[tbh!]\begin{center}
1383: $a)$\hspace{8cm}$b)$
1384: \includegraphics[width=3.2in,angle=0]{plots/LQC-rho-z1.eps}
1385: \includegraphics[width=3.2in,angle=0]{plots/LQC-rho-z2.eps}
1386: \caption{A detailed picture of the comparizon of
1387: $\expect{\rho_{\phi}}$ against the $\rho_{\phi}(\phi)$ trajectories
1388: presented in fig.~\ref{fig:vr-traj}b is shown near the bounce $(a)$ and
1389: recollapse $(b)$ points respectively.}
1390: \label{fig:rho-zoom}
1391: \end{center}\end{figure}
1392:
1393: % DISCUSSION
1394:
1395: The results listed above show that the picture based on the analysis of the
1396: previous models is valid here as well. Similarly to that cases, the
1397: correction due to the discreteness of geometry cause gravity to become
1398: repulsive at large energy densities and, in particular, force the
1399: universe to bounce when the energy density reaches a critical value. This
1400: indicates that $\rho_c$ may be a fundamental quantity, independent on
1401: the matter content at least in isotropic cases. Furthermore, the
1402: states remain sharply peaked even in regions where quantum-geometric
1403: effects dominate the dynamics, where in principle one expects to
1404: loose the semiclassicality. The dynamics itself can be well approximated by
1405: an effective Friedmann equation (see Appendix \ref{sec:eff})
1406: \begin{equation}
1407: H^2 = \fracs{8\pi G}{3}\rho(1-\rho/\rho_c) \ ,
1408: \end{equation}
1409: where $H$ is a Hubble rate and $\rho$ is a total energy density.
1410:
1411: % dispersion
1412: The agreement between the quantum evolution and the effective one brings out
1413: another issue: since the spectrum of $\Theta$ is not exactly uniform, the
1414: states are not exactly periodic and a spread increase can be observed
1415: between cycles. This leads ultimately to the loss of semiclassicality.
1416: This in turn raises the question about the size of time interval in which
1417: the state remains sharply peaked.
1418:
1419: To answer this question, we analyzed the spread increase within one cycle of
1420: evolution. It can be estimated via the heuristic methods
1421: described in section \ref{sec:heuDisp} and turns out to be much smaller than in
1422: $k=1$ case. For example, when $\Lambda \approx -10^{-120}$, a
1423: universe peaked about $p^{\star}_{\phi}$ large enough for it to grow
1424: to megaparsec size, and with relative dispersions in $p_{\phi}$ and
1425: $v$ of the same order, will need at least $10^{70}$ cycles for the
1426: relative dispersion to double. The number of cycles needed to grow
1427: to a considerably large value (say $10^{-6}$) is correspondingly larger.
1428:
1429: For small values of the momentum (that is $p_{\phi}^{\star}
1430: \leq 10^3\hbar$), we were able to confirm the heuristics numerically.
1431: Also, since for larger momenta the states become more and more
1432: semiclassical, we expect the estimate to become more accurate
1433: there. The result is however far from general, as numerical tests were
1434: done for a specific family of states only, thus (as it was discussed
1435: in \cite{mb-L}) do not allow us to exclude the situation, where some
1436: specific example of state violates the bound. On the other hand, the
1437: proposed estimate is based on the properties of the spectrum of
1438: $\Theta$, thus we expect that a bound of at least a similar order
1439: should hold in general. Such situation happened for example in the
1440: $\Lambda=0$ case \cite{cs-rec}, where it was possible to find (in the
1441: context of sLQC) an analogous bound satisfied by all the states
1442: which admit semiclassical epoch (see section \ref{sec:intro}) in their
1443: history. A similar bound was next derived in exact LQC
1444: \cite{kp-rec}. More precise statements regarding model considered here
1445: will however require further work.\\
1446:
1447: \noindent{{\bf Acknowledgments}} We would like to thank Abhay Ashtekar,
1448: Wojciech Kami\'nski and Parampreet Singh for extensive discussions and
1449: helpful comments. We also profitted from discussions with Martin
1450: Bojowald and Jerzy Lewandowski. This work was supported in part by the
1451: National Science Foundation (NSF) grant PHY-0456913 and the Eberly
1452: research funds. TP acknowledges financial aid provided by the I3P
1453: framework of CSIC and the European Social Fund. EB acknowledges the
1454: support of the Center for Gravitational Wave Physics, funded by the
1455: National Science Foundation under Cooperative Agreement PHY-0114375.
1456:
1457: \appendix
1458:
1459: \section{Anti-symmetric sector of LQC}\label{sec:antisymm}
1460:
1461: Due to the lack of a parity violating interaction in the model considered in
1462: this article, the change in the triad orientation is a large gauge symmetry.
1463: This allowed us to restrict the physical Hilbert space to the subspace
1464: of states invariant with respect to the reflection in $v$
1465: corresponding to this orientation change -- the symmetric sector. In
1466: principle, however, we could choose instead the space of states which
1467: are antisymmetric under considered transformation. There is no
1468: physical reason to favor one of these two choices over the other. This
1469: raised a concerns on whether the results of LQC are tied to the
1470: selection of symmetric sector and whether they will still hold in the
1471: antisymmetric one. We address these concerns here by repeating the
1472: constructions of section \ref{sec:LQC}, this time building the
1473: physical Hilbert space out of antisymmetric states.
1474:
1475: First, following section \ref{sec:LQC} we divide the kinematical
1476: Hilbert space $\Hilkg$ onto superselection sectors, i.e.~the spaces $\Hilkge$
1477: of functions supported on lattices $\lat_{\varepsilon}$. The results
1478: of \cite{klp-nL} (self-adjointness of $\Theta$ and discreteness of its
1479: spectrum on each of these spaces) were derived without any symmetry
1480: assumption, thus they hold also in our case. Furthermore, as we will show
1481: below, the spectrum is non-degenerate also when we restrict the space
1482: of eigenfunctions to the antisymmetric ones. In consequence, we can
1483: construct the physical Hilbert space as specified in \eqref{eq:Psi},
1484: but by imposing on the relevant eigenfunctions $e_n^a$ the condition
1485: $e^a_{n}(v) = -e^a_{n}(-v)$ instead of the symmetry requirement.
1486:
1487: To check the effect of the above modification on the dynamics, we have to
1488: examine how it changes the exact form of $e_n$. That, in turn, depends
1489: on the value of the superselection sector label $\varepsilon$.
1490: \begin{itemize}
1491: \item For $\epsilon\neq 0,2$ (generic lattices), the symmetric
1492: eigenfunction on $\lat_{\varepsilon}$ is completely determined (see
1493: discussion in section \ref{sec:num-eig}) by its restriction to the
1494: lattice $\lat_{+|\varepsilon|}$, with the remaining part supported
1495: on $\lat_{-|\varepsilon|}$ determined via a symmetry
1496: relation. Furthermore, symmetry does not impose any constraint on
1497: the part supported on $\lat_{+|\varepsilon|}$ itself and we can
1498: complete it to the antisymmetric eigenfunction by simply acting with
1499: $-\Pi$ on it. In consequence, there exists a $1-1$ correspondence between
1500: these two types of eigenfunctions. Namely, each antisymmetric
1501: eigenfunction $\psi^a$ is related to the symmetric one $\psi$ via:
1502: \begin{equation}\label{eq:sa-eg}
1503: \psi^a(v)\ =\ \begin{cases} \psi(v) & v\in\lat_{+|\varepsilon|} \\
1504: -\psi(v) & v\in\lat_{-|\varepsilon|} \end{cases} \ ,
1505: \end{equation}
1506: This implies that, in the antisymmetric sector, the spectrum of
1507: $\Theta$ is the same as in the symmetric one.
1508: \item When $\epsilon=2$, the situation is similar to the generic case.
1509: The solutions to \eqref{eq:eigen} on two sublattices
1510: $\lat_{\varepsilon}\cap\re^+$ and $\lat_{\varepsilon}\cap\re^-$
1511: are independent, thus each eigenfunction is again determined by
1512: its restriction to $\lat_{\varepsilon}\cap\re^+$. In consequence,
1513: we again have the $1-1$ correspondence between the symmetric and
1514: antisymmetric eigenfunctions
1515: \begin{equation}\label{eq:sa-e2}
1516: \psi^a(v)\ =\ \begin{cases} \psi(v) & v>0 \\
1517: -\psi(v) & v<0 \end{cases} \ ,
1518: \end{equation}
1519: and the spectra of $\Theta$ in both sectors are identical.
1520: \item The case when $\epsilon=0$ requires a bit more care. In section
1521: \ref{sec:num-eig}, the symmetry assumption imposed on solutions to
1522: \eqref{eq:eigen} an additional constraint, allowing to determine
1523: $\psi(4)$ for known $\psi(0)$. Antisymmetry replaces this
1524: constraint by a different one: $\psi^a(0)=0$. Therefore the whole
1525: procedure of identifying the normalizable eigenfunctions has to be
1526: redone. We can however apply exactly the same procedure as in
1527: section \ref{sec:num-eig}. The results are as follows:
1528: \begin{itemize}
1529: \item The qualitative features of the eigenfunctions remain the
1530: same. In particular, we can still distinguish the same 5 zones of
1531: exponential/oscillatory behavior (see fig.\ref{fig:antisymm}a).
1532: Their boundaries $v_B$, $v_R$ are exactly the same as in the
1533: symmetric case.
1534: \item The spectrum of $\Theta$ in the antisymmetric sector is
1535: different than in the symmetric case, however the eigenvalues of
1536: one sector approach the ones of the other very quickly (see
1537: fig.\ref{fig:antisymm}b). In consequence, the separation
1538: between the eigenvalues approaches, as $\omega\to\infty$, the same
1539: limit shown in fig.~\ref{fig:dw}b. The rate of approach to
1540: this limit also remains the same.
1541: \end{itemize}
1542: \end{itemize}
1543:
1544: The similarity between eigenfunctions of the two considered sectors
1545: implies an analogous similarity between the physical states. In particular,
1546: for $\varepsilon\neq 0$, if the eigenfunctions satisfy equations
1547: (\ref{eq:sa-eg}-\ref{eq:sa-e2}), so will the wave
1548: functions. Then if we take two physical states, a symmetric and
1549: an antisymmetric one with the same spectral profile $\tilde{\Psi}_n$, the
1550: expectation values of (all the powers of) the observables defined in
1551: section \ref{sec:lqc-obs} will be {\it exactly the same} for both of
1552: them.
1553:
1554: For $\varepsilon=0$, due to the slight difference in the spectrum, we have to
1555: repeat the analysis of \ref{sec:num-state}. But again the numerical
1556: checks reveal no measurable deviations from the results obtained in the symmetric
1557: sector.
1558:
1559: In summary the results obtained for both $\varepsilon\neq 0$ and
1560: $\varepsilon=0$ show that working with the antisymmetric sector instead of
1561: the symmetric one does not produce any qualitative changes or (apart from
1562: a slightly different spectrum of $\Theta$ in $\varepsilon=0$ case)
1563: any measurable modifications to the physics predicted by the model.
1564:
1565: \begin{figure}[tbh!]\begin{center}
1566: $a)$\hspace{8cm}$b)$
1567: \includegraphics[width=3.2in,angle=0]{plots/eigenval-anti.eps}
1568: \includegraphics[width=3.2in,angle=0]{plots/eigenfs-anti.eps}
1569: \caption{$a)$ A set of lowest ($\omega<44$) spectrum elements
1570: of $\Theta$ for the symmetric and antisymmetric sector (with
1571: $\Lambda=-1$) is
1572: shown with respect to $\pm|\varepsilon|$. The green x-es
1573: represent the eigenvalues corresponding to the cases where the relation
1574: between symmetric and antisymmetric eigenfunctions is given by
1575: (\ref{eq:sa-eg}, \ref{eq:sa-e2}) (denoted as {\it generic}). The
1576: red crosses and blue stars represent the eigenvalues of, respectively,
1577: the symmetric and antisymmetric eigenfunctions on the lattice
1578: $\lat_{\varepsilon=0}$. The
1579: antisymmetric eigenfunctions $e^a_{0}$ to $e^a_{4}$ corresponding
1580: to the eigenvalues shown in $a)$ are presented in $b)$. For clarity,
1581: they are plotted on the $v>0$ semiaxis only.}
1582: \label{fig:antisymm}
1583: \end{center}\end{figure}
1584:
1585:
1586: \section{Heuristic description}\label{sec:heu}
1587:
1588: In this sections we discuss some issues related to the heuristic method
1589: for the description of the quantum evolution. We divide its content
1590: into two parts, dedicated respectvely to the effective classical
1591: dynamics and the estimate of the dispersion growth during the
1592: evolution of the semiclassical state.
1593:
1594:
1595: \subsection{Effective dynamics}\label{sec:eff}
1596:
1597: The numerical tests described in the main body of the paper have shown that
1598: if a state is semiclassical at some epoch, it will remains so for
1599: a large fraction of the evolution (i.e., a large number of cycles of
1600: bounces and recollapses).
1601: In particular it remains sharply peaked even in the regions where the quantum
1602: gravity corrections modify the dynamics. This indicates the existence
1603: of a (n effective) classical theory whose predictions well agree with
1604: those of LQC.
1605:
1606: At the rigorous level, such theory was derived for $\Lambda=0$
1607: \cite{VT} with the use of the geometric formulation of quantum mechanics
1608: \cite{GQM}. Up to the second order quantum corrections (remaining
1609: always small during the evolution), its results confirm the
1610: predictions of the classical effective dynamics proposed earlier
1611: \cite{pv,aps-imp}, derived heuristically by replacing the connection $c$ in
1612: classical Hamiltonian by $\sin(\bar{\mu}c)/\bar{\mu}$. Here we apply
1613: this heuristic method to the system with a cosmological constant
1614: considered in the main body of the paper. An analogous derivation of
1615: the effective dynamics (to the level of quadratures) and the analysis of
1616: the resulting trajectories was done in \cite{eff}, however the
1617: trajectory parametrization used there makes the direct comparizon with
1618: the results of quantum evolution difficult.
1619:
1620: The classical Hamiltonian of the system is related to the constraint
1621: (\ref{eq:classHgrav}, \ref{eq:classHcompl}) via $\Heff = C/(16\pi G)$.
1622: Application of the rule $c \to \sin(\bar{\mu}c)/\bar{\mu}$ yields the
1623: result
1624: \begin{equation}\label{eq:Heff}
1625: \Heff\ =\
1626: -\frac{3}{8\pi G\gamma^2\bar{\mu}^2}|p|^{\frac{1}{2}}\sin^2(\bar{\mu}c)
1627: + \frac{1}{2}\frac{p^{2}_{\phi}}{|p|^{\frac{3}{2}}}
1628: + \frac{p^{\frac{3}{2}}}{16\pi G} \Lambda \ .
1629: \end{equation}
1630: Hamilton's equations $\dot{v} = \{v,\Heff\}$ and $\dot{\phi} =
1631: \{\phi,\Heff\}$ are identical to the $\Lambda=0$ case. Written in terms
1632: of $v$ they are respectively
1633: \begin{subequations}\label{eq:Ham}\begin{align}
1634: \label{eq:vdot}
1635: \dot{v}\ &=\ \frac{3v}{\sqrt{2\sqrt{3}\pi\gamma\lPl^2}}
1636: \sin(\bar{\mu}c) \cos(\bar{\mu}c) \ , &
1637: \dot{\phi}\ &=\ \left(\frac{6}{8\pi\gamma\lPl^2}\right)^{\frac{3}{2}}
1638: \frac{K}{|v|}p_{\phi} \ .
1639: \tag{\ref{eq:Ham}}\end{align}\end{subequations}
1640: Taking the square of (\ref{eq:Ham}a) and supplying $\sin(\bar{\mu}c)$
1641: via \eqref{eq:Heff}, we arrive to an analog of Friedmann equation:
1642: \begin{equation}\label{eq:FriedEff}
1643: H^2\ :=\ \left(\frac{\dot{v}}{3v}\right)^2\ =\
1644: \frac{8\pi G}{3} \rho \left( 1-\frac{\rho}{\rho_c} \right) \ ,
1645: \end{equation}
1646: where $\rho$ and $\rho_c$ are the total matter energy density and the critical
1647: energy density found in \cite{aps-imp}
1648: \begin{subequations}\label{eq:RhoRho}\begin{align}
1649: \rho\ &:=\ \rho_{\phi}+\frac{\Lambda}{8\pi G} \ , &
1650: \rho_{\phi}\ &:=\ \frac{p^2_{\phi}}{2p^3} \ , &
1651: \rho_c\ &:=\ \frac{\sqrt{3}}{16\pi^2\gamma^3 G^2\hbar} \ .
1652: \tag{\ref{eq:RhoRho}}\end{align}\end{subequations}
1653: Applying (\ref{eq:Ham}b), we can rewrite the resulting Friedmann
1654: equation in the form involving $v$ and $\phi$ only
1655: \begin{equation}\label{eq:Fried1}
1656: v_{\phi}\ =\ \pm\, v\,\sqrt{12\pi G} \left[ \frac{\rho}{\rho_{\phi}}
1657: \left(1-\frac{\rho}{\rho_c}\right)\right]^{\frac{1}{2}} \ .
1658: \end{equation}
1659: The sign in front of the right hand side of the above equation depends on
1660: the evolution epoch, and in particular changes during
1661: recollapse. Therefore, it is convenient to rewrite \eqref{eq:Fried1} in
1662: the second order form (obtained by differentiating it):
1663: \begin{equation}\label{eq:Fried2}
1664: v_{\phi\phi}\ =\ 12\pi G\, v
1665: \left[ \left(\frac{2\rho}{\rho_{\phi}}-1\right)
1666: \left(1-\frac{\rho}{\rho_c}\right)
1667: + \frac{\rho}{\rho_c} \right] \ .
1668: \end{equation}
1669:
1670: % numerical solutions
1671: To compare the results of the numerical evolution, we integrated
1672: equation \eqref{eq:Fried2} numerically using a fifth-order adaptive
1673: Runge-Kutta method (known as {\it RK45}). The initial value $\dot{v}$,
1674: needed to complete the initial data specification, was calculated via
1675: \eqref{eq:Fried1}.
1676:
1677: An example of the comparison results is shown in
1678: figs.~\ref{fig:vr-traj},~\ref{fig:v-zoom},~\ref{fig:rho-zoom}. The
1679: trajectories agree with the results of the quantum evolution everywhere.
1680: The differences between them are much smaller than the spreads of the wave
1681: packets even near the bounce.
1682:
1683:
1684: \subsection{Bound on the dispersion growth}\label{sec:heuDisp}
1685:
1686: The analysis of section \ref{sec:num} has shown that the states that are
1687: semiclassical at a given initial time $\phi_o$ remain so for many cycles
1688: of bounces and recollapses. However, due to non-uniformity of the
1689: spectrum of the $\Theta$ operator, the initially coherent wave packet slowly
1690: spreads out. Here we derive an upper bound on this spread growth using
1691: some heuristic estimates and applying the knowledge about the spectrum
1692: of $\Theta$ presented in section \ref{sec:num-eig}.
1693:
1694: To start with, let us note that for large $p_{\phi}=\hbar\omega$ the
1695: distance between neighboring eigenvalues is almost constant
1696: $\omega_{n+1}-\omega_n \approx \Delta\omega$ (see \eqref{eq:dw}). In
1697: consequence, the wave function is almost periodic in $\phi$, with an
1698: approximate period equal to
1699: \begin{equation}\label{eq:tw}
1700: T\ \approx\ \frac{2\pi}{\Delta\omega} \ .
1701: \end{equation}
1702: Now, if we consider two classical (effective) trajectories
1703: corresponding to $p_{\phi}$ equal respectively to $\hbar\omega$ and
1704: $\hbar\omega+\delta\omega$, the difference between periods is
1705: determined by the corrections to the uniformity of
1706: $\Delta\omega_n$. They are in turn bounded by the function
1707: $A\,(\Delta\omega)^2\,\omega^{-2}$ (see \eqref{eq:A-bound}). Applying
1708: this bound to \eqref{eq:tw} (i.e. taking $\Delta\omega_n =
1709: \Delta\omega(1+A\omega^{-2})$) and neglecting terms of higher order in
1710: $\delta\omega$, we obtain the following difference in $T$:
1711: \begin{equation}
1712: \delta T\ \approx\ \frac{4\pi A}{\omega^3}\delta\omega \ ,
1713: \end{equation}
1714: which can be now used to estimate the growth of $\Delta v/v$ within one
1715: cycle. To do so, we note that, since the cosmological constant term acts
1716: like a positive $v^2$ potential, the speed $v_{\phi}$ is bounded from
1717: above by the speed $v^o_{\phi}$ of a classical universe with $\Lambda=0$
1718: \begin{equation}
1719: |v_{\phi}|\ \leq\ |v^o_{\phi}|\ :=\ \sqrt{12\pi G}|v| \ .
1720: \end{equation}
1721: In consequence:
1722: \begin{equation}\label{eq:ddv-bound}
1723: \frac{\delta v}{v}\ \leq\ 8\sqrt{3}\pi^{\frac{3}{2}}A\,
1724: \frac{1}{\omega^2}\,
1725: \frac{\delta\omega}{\omega} \ .
1726: \end{equation}
1727:
1728: In order to arrive to this bound, we used some heuristic methods that need to
1729: be confirmed using numerics. Unfortunately, due to the extremely small
1730: values of $\delta v/v$, we were able to check \eqref{eq:ddv-bound} only
1731: for small values of the frequency, $\omega\leq 10^3$. To do so, we
1732: calculated the semiclassical states in two intervals of $\phi$
1733: separated by a large ($>100$) distance in $\phi$ and compared the
1734: difference between the maximal relative dispersion in $v$ observed
1735: within one cycle in both of the chosen intervals. To compute the wave
1736: functions, we used a direct summation method specified in section
1737: \ref{sec:num-state}. Within the checked range $500\leq\omega\leq 1000$, the
1738: bound was satisfied.
1739:
1740:
1741:
1742: \begin{thebibliography}{99}
1743: \bibitem{lqc} M.~Bojowald, {Loop Quantum Cosmology}, Living
1744: Rev.Rel. \textbf{8}, 11 (2005), \texttt{arXiv: gr-qc/0601085};\\
1745: A.~Ashtekar, {An Introduction to Loop Quantum Gravity Through
1746: Cosmology}, Nuovo Cim. \textbf{122B}, 135-155 (2007),
1747: \texttt{arXiv:gr-qc/0702030}.
1748: \bibitem{lqg}
1749: C.~Rovelli, {\em Quantum Gravity} (CUP, Cambridge, 2004);\\
1750: A.~Ashtekar and J.~Lewandowski, {Background Independent Quantum
1751: Gravity: A Status Report}, Class.Quant.Grav. \textbf{21}, R53
1752: (2004), \texttt{arXiv:gr-qc/0404018};\\
1753: T.~Thiemann, {\em Introduction to Modern Canonical
1754: Quantum General Relativity} (CUP, Cambridge, 2007).
1755: \bibitem{aps} A.~Ashtekar, T.~Pawlowski and P.~Singh, {Quantum nature
1756: of the big bang}, Phys.Rev.Lett. \textbf{96}, 141301 (2006),
1757: \texttt{arXiv:gr-qc/0602086}.
1758: \bibitem{aps-old} A.~Ashtekar, T.~Pawlowski and P.~Singh, {Quantum nature
1759: of the big bang: An analytical and numerical investigation},
1760: Phys.Rev. {\bf D73}, 124038 (2006), \texttt{arXiv:gr-qc/0604013}.
1761: \bibitem{aps-imp} A.~Ashtekar, T.~Pawlowski and P.~Singh, {Quantum nature
1762: of the big bang: Improved dynamics}, Phys.Rev. {\bf D74}, 084003
1763: (2006), \texttt{arXiv:gr-qc/0607039}.
1764: \bibitem{pv} P.~Singh and K.~Vandersloot, {Non-Singular Bouncing
1765: Universes in Loop Quantum Cosmology}, Phys.Rev. \textbf{D72}
1766: 084004, (2005), \texttt{arXiv:gr-qc/0606032}.
1767: \bibitem{apsv} A.~Ashtekar, T.~Pawlowski, P.~Singh and
1768: K.~Vandersloot, {Loop quantum cosmology of k=1 FRW models},
1769: Phys.Rev. {\bf D75}, 024035 (2007), \texttt{arXiv:gr-qc/0612104}.
1770: \bibitem{klp} L.~Szulc, W.~Kaminski and J.~Lewandowski,
1771: {Closed FRW model in Loop Quantum Cosmology}, Class.Quant.Grav.
1772: \textbf{24}, 2621-2635 (2007), \texttt{arXiv:gr-qc/0612101}.
1773: \bibitem{klp-nL} W.~Kaminski and J.~Lewandowski,
1774: {The flat FRW model in LQC: the self-adjointness}, (2007),
1775: \texttt{arXiv:0709.3120}.
1776: \bibitem{acs} A.~Ashtekar, A.~Corichi and P.~Singh, {Robustness of
1777: key features of loop quantum cosmology}, (2007),
1778: \texttt{arXiv:0710.3565}.
1779: \bibitem{cs-rec} A.~Corichi, P.~Singh, {Quantum bounce and cosmic
1780: recall}, (2007), \texttt{arXiv:0710.4543}.
1781: \bibitem{chv} D.~W.~Chiou, {Loop Quantum Cosmology in Bianchi Type I
1782: Models: Analytical Investigation}, Phys.Rev. \textbf{D75}, 024029
1783: (2007), \texttt{arXiv:gr-qc/0609029};\\
1784: D.~W.~Chiou, {Effective Dynamics for the Cosmological Bounces in
1785: Bianchi Type I Loop Quantum Cosmology}, (2007),
1786: \texttt{arXiv:gr-qc/0703010};\\
1787: D.~W.~Chiou and K.~Vandersloot, {The behavior of non-linear
1788: anisotropies in bouncing Bianchi I models of loop quantum
1789: cosmology}, Phys.Rev. \textbf{D76}, 084015 (2007),
1790: \texttt{arXiv:0707.2548};\\
1791: D.~W.~Chiou, {Effective Dynamics, Big Bounces and Scaling Symmetry
1792: in Bianchi Type I Loop Quantum Cosmology},
1793: Phys.Rev. \textbf{D76},
1794: 124037 (2007), \texttt{arXiv:0710.0416};\\
1795: L.~Szulc, {Loop Quantum Cosmology of Diagonal Bianchi Type I
1796: model: simplified theory}, (2008), \texttt{arXiv:0803.3559}.
1797: \bibitem{gp} M.~Campiglia, R.~Gambini and J.~Pullin, {Loop
1798: quantization of spherically symmetric midi-superspaces},
1799: Class.Quant.Grav. \textbf{24}, 3649-3672 (2007),
1800: \texttt{arXiv:gr-qc/0703135}.
1801: \bibitem{gm-letter} M.~Mart\'in-Benito, L.~J.~Garay and G.~A.~Mena
1802: Marug\'an, {Hybrid Quantum Gowdy Cosmology: Combining Loop and Fock
1803: Quantization}, (2008), \texttt{arXiv:0804.1098}.
1804: \bibitem{eff} J.~Mielczarek, T.~Stachowiak, M.~Szyd{\l}owski, {Exact
1805: solutions for Big Bounce in loop quantum cosmology}, (2008),
1806: \texttt{arXiv:0801.0502}.
1807: \bibitem{abl} A.~Ashtekar, M.~Bojowald, J.~Lewandowski,
1808: {Mathematical structure of loop quantum cosmology}, Adv.Theo.%
1809: Math.Phys. \textbf{7}, 233-268 (2003), \texttt{arXiv:gr-qc/0304074}.
1810: \bibitem{lqc-MB}
1811: M.~Bojowald, {Absence of singularity in loop quantum cosmology},
1812: Phys.Rev.Lett. \textbf{86}, 5227-5230 (2001),
1813: \texttt{arXiv:gr-qc/0102069};\\
1814: M.~Bojowald, {Isotropic loop quantum cosmology},
1815: Class.Quant.Grav. \textbf{19}, 2717-2741 (2002),
1816: \texttt{arXiv:gr-qc/0202077}.
1817: \bibitem{ThTrick} T.~Thiemann, Anomaly-free formulation of
1818: non-perturbative, four-dimensional Lorentzian quantum gravity,
1819: {Phys.Lett.} \textbf{B380}, 257-264 (1998),
1820: \texttt{gr-qc/9606088};\\
1821: T.~Thiemann, Quantum spin dynamics (QSD),
1822: {Class.Quant.Grav.} \textbf{15}, 839-873 (1998),
1823: \texttt{gr-qc/9606089};\\
1824: T.~Thiemann, QSD V : Quantum gravity as the natural regulator of
1825: matter quantum field theories, {Class.Quant.Grav.} \textbf{15},
1826: 1281-1314 (1998), \texttt{gr-qc/9705019}.
1827: \bibitem{gave}
1828: D.~Marolf, Refined algebraic quantization: Systems with a single
1829: constraint, (1995), \texttt{arXiv:gr-qc/9508015};\\
1830: D.~Marolf, {Quantum observables and recollapsing dynamics},
1831: Class.Quant.Grav. {\bf 12}, 1199-1220 (1995),
1832: \texttt{arXiv:gr-qc/9404053};\\
1833: D.~Marolf, {Observables and a Hilbert space for Bianchi IX},
1834: Class.Quant.Grav. {\bf 12}, 1441-1454 (1995),
1835: \texttt{arXiv:gr-qc/9409049};\\
1836: D.~Marolf, {Almost ideal clocks in quantum cosmology: A brief
1837: derivation of time}, Class.Quant.Grav. {\bf 12}, 2469-2486 (1995),
1838: \texttt{arXiv:gr-qc/9412016};\\
1839: A.~Ashtekar, J.~Lewandowski, D.~Marolf,
1840: J.~Mour\~ao and T.~Thiemann, Quantization of diffeomorphism
1841: invariant theories of connections with local degrees of freedom,
1842: {J.Math.Phys.} \textbf{36}, 6456-6493 (1995),
1843: \texttt{arXiv:gr-qc/9504018}.
1844: \bibitem{BesK} A.~Gil, J.~Segura and N.~M.~Temme, Evaluation of the
1845: modified Bessel function of the third kind of imaginary orders, {
1846: J.Comput.Phys.} {\bf 175}, 398-411 (2002);\\
1847: A.~Gil, J.~Segura and N.~M.~Temme, Computation of the modified Bessel
1848: function of the third kind of imaginary orders: uniform Airy-type
1849: asymptotic expansion, {J.Comput.App.Math.} {\bf 153}, 225-234
1850: (2003);\\
1851: A.~Gil, J.~Segura and N.~M.~Temme,
1852: Computing solutions of the modified Bessel differential equation for
1853: imaginary orders and positive arguments,
1854: {ACM Trans.Math.Soft.} {\bf 30}, 145-158 (2004),
1855: \texttt{arXiv:math/0401128};\\
1856: A.~Gil, J.~Segura and N.~M.~Temme,
1857: Algorithm 831: Modified Bessel functions of imaginary order and
1858: positive argument, {ACM Trans.Math.Soft.} {\bf 30}, 159-164
1859: (2004), \texttt{arXiv:cs/0401008}.
1860: \bibitem{mb-L} M.~Bojowald, {Harmonic cosmology: How much can we
1861: know about a universe before the big bang?}, (2007),
1862: \texttt{arXiv:0710.4919v1};\\
1863: M.~Bojowald, {Quantum nature of cosmological
1864: bounces}, (2008), \texttt{arXiv:0801.4001}.
1865: \bibitem{kp-rec} W.~Kami\'nski and T.~Pawlowski, (2008),
1866: {\it in preparation}
1867: \bibitem{VT} V.~Taveras, IGPG preprint, (2006).
1868: \bibitem{GQM} A.~Ashtekar and T.~Schilling, Geometrical formulation of
1869: quantum mechanics, In: \textit{On Einstein's path}, A.~Harvery, ed
1870: (Springer-Verlag, New York, 1998), \texttt{arXiv:gr-qc/9706069}.
1871: T.~Schilling, Geometry of quantum mechanics, Ph.D. Dissertation,
1872: Penn State (1996), \texttt{%
1873: http://cgpg.gravity.psu.edu/archives/thesis/index.shtml}
1874: \end{thebibliography}
1875:
1876: \end{document}
1877: