1: % Artur's final corrections.
2:
3: \documentclass[12pt,pre,eqsecnum,aps,epsfig,floats]{revtex4}
4:
5: \bibliographystyle{prsty}
6:
7: \usepackage{graphics}
8: \usepackage{epsfig}
9: \usepackage{amsfonts}
10: \usepackage{amsmath}
11: \usepackage{theorem}
12:
13: \def\mybox{\,\vrule height7pt width6.2pt depth-1pt}
14:
15: \newcommand{\one}{\mbox{\tt 1}\hspace{-0.057 in}\mbox{\tt l}}
16:
17: \newcommand{\Ket}[1]{|#1\rangle}
18:
19: \newcommand{\ba}{\mbox{\boldmath $a$}}
20: \newcommand{\g}{\mbox{\boldmath $g$}}
21: \newcommand{\f}{\mbox{\boldmath $f$}}
22: \newcommand{\gs}{\mbox{\boldmath $\scriptstyle g$}}
23: \newcommand{\bu}{\mbox{\boldmath $u$}}
24: \newcommand{\us}{\mbox{\boldmath $\scriptstyle u$}}
25: \newcommand{\bv}{\mbox{\boldmath $v$}}
26: \newcommand{\vs}{\mbox{\boldmath $\scriptstyle v$}}
27: \newcommand{\w}{\mbox{\boldmath $w$}}
28: \newcommand{\ws}{\mbox{\boldmath $\scriptstyle w$}}
29: \newcommand{\x}{\mbox{\boldmath $x$}}
30: \newcommand{\xs}{\mbox{\boldmath $\scriptstyle x$}}
31: \newcommand{\y}{\mbox{\boldmath $y$}}
32: \newcommand{\ys}{\mbox{\boldmath $\scriptstyle y$}}
33: \newcommand{\z}{\mbox{\boldmath $z$}}
34: \newcommand{\zs}{\mbox{\boldmath $\scriptstyle z$}}
35: \newcommand{\as}{\mbox{\boldmath $\scriptstyle a$}}
36: \newcommand{\at}{\mbox{\boldmath $a$}}
37: \newcommand{\bs}{\mbox{\boldmath $\scriptstyle b$}}
38: \newcommand{\bt}{\mbox{\boldmath $b$}}
39: \newcommand{\xibold}{\mbox{\boldmath $\xi$}}
40: \newcommand{\xibolds}{\mbox{\boldmath $\scriptstyle \xi$}}
41: \newcommand{\gammas}{\mbox{$\scriptstyle \gamma$}}
42: \newcommand{\boldarrow}{\mbox{\boldmath $\swarrow$}}
43:
44: \newcommand{\Tr}{\mbox{\rm\small Tr\ }}
45: \newcommand{\balpha}{\mbox{\boldmath $\alpha$}}
46: \newcommand{\balphas}{\mbox{\boldmath $\scriptstyle \alpha$}}
47: \newcommand{\bbeta}{\mbox{\boldmath $\beta$}}
48: \newcommand{\bbetas}{\mbox{\boldmath $\scriptstyle \beta$}}
49:
50: \newcommand{\bgamma}{\mbox{\boldmath $\gamma$}}
51: \newcommand{\bdelta}{\mbox{\boldmath $\delta$}}
52: \newcommand{\bsigma}{\mbox{\boldmath $\sigma$}}
53: \newcommand{\bxi}{\mbox{\boldmath $\xi$}}
54: \newcommand{\bxis}{\mbox{\boldmath $\scriptstyle\xi$}}
55: \newcommand{\boldeta}{\mbox{\boldmath $\eta$}}
56: \newcommand{\boldetas}{\mbox{\boldmath $\scriptstyle\eta$}}
57: \newcommand{\bzeta}{\mbox{\boldmath $\zeta$}}
58: \newcommand{\bzetas}{\mbox{\boldmath $\scriptstyle\zeta$}}
59:
60: \newcommand{\bnabla}{\mbox{\boldmath $\nabla$}}
61:
62: \theoremstyle{break}
63:
64: \newtheorem{definition}{Definition}
65: \newtheorem{remark}{Remark}
66: \newtheorem{lemma}{Lemma}
67: \newtheorem{theorem}{Theorem}
68: \newtheorem{ThProof}{Proof of Theorem}
69: \newtheorem{example}{Example}
70: \newtheorem{identity}{Identity}
71:
72: \begin{document}
73:
74:
75: \title{Classical predictability and coarse-grained
76: evolution of the quantum baker's map}
77: \author{Artur Scherer}
78: \author{Andrei N. Soklakov}
79: \author{R\"udiger Schack}
80: \affiliation{Department of Mathematics,
81: Royal Holloway, University of London,
82: Egham, Surrey TW20 0EX, UK}
83:
84: \date{\today}
85:
86: \begin{abstract}
87: We investigate how classical predictability of the coarse-grained evolution
88: of the quantum baker's map depends on the character of the coarse-graining.
89: Our analysis extends earlier work by Brun and Hartle [Phys.\ Rev.\ D {\bf
90: 60}, 123503 (1999)] to the case of a chaotic map. To quantify
91: predictability, we compare the rate of entropy increase for a family of
92: coarse-grainings in the decoherent histories formalism. We find that the
93: rate of entropy increase is dominated by the number of scales characterising
94: the coarse-graining.
95: \end{abstract}
96:
97: \maketitle
98:
99: \section{Introduction}
100:
101:
102: The concept of coarse-graining plays an important role in the emergence of
103: classical evolution from the fundamental quantum-mechanical equations of
104: motion~\cite{Gell-MannHartle1993,BrunHartle1999-PRD}. The form of the
105: effective classical equations of motion is as much influenced by the character
106: of the coarse-graining as by the fundamental quantum-mechanical equations of
107: motion themselves. A systematic way to study coarse-grained quantum evolution
108: is provided by the decoherent histories
109: formalism~\cite{Griffiths1984,Omnes1988,Gell-MannHartle1990,DowkerHalliwell1992,Gell-MannHartle1993}.
110: Within this approach to quantum theory a quantum mechanical system is said to
111: exhibit classical behaviour when histories with correlations in time that are
112: implied by classical deterministic laws have high
113: probability~\cite{Gell-MannHartle1993,BrunHartle1999-PRD}.
114:
115: Coarse-grained descriptions are also used in classical physics to reduce the
116: number of variables when the number of degrees of freedom is large. This leads
117: to effective equations of motion for the coarse-grained variables. The
118: character of the coarse-graining is important here. Although a given physical
119: system may be described by many alternative sets of coarse-grained variables,
120: some coarse-grained descriptions are more useful for prediction than others.
121: For a practical set of coarse-grained variables, the observables of interest
122: should be simple and slowly varying functions.
123:
124: In quantum theory, the nonuniqueness of the coarse-graining procedure motivates
125: this question: what distinguishes coarse-grainings leading to predictable,
126: deterministic effective classical evolution from other coarse-grainings? In
127: general, arbitrarily many sets of alternative coarse-grained histories
128: decohere and so can be assigned probabilities. Moreover, two such decoherent
129: sets of histories are in general mutually incompatible. Which of these many
130: possible coarse-grainings lead to predictable evolution of the coarse-grained
131: variables, i.e., useful regularities in time governed by effective,
132: phenomenological equations of motion?
133:
134: These questions have been addressed by Brun and Hartle in
135: Ref.~\cite{BrunHartle1999-PRD}, where they investigate the origin of classical
136: predictability by considering the simplest linear system with a continuum
137: description---the linear one-dimensional harmonic chain regarded as a closed
138: quantum mechanical system. In their analysis a chain of $\mathcal{N}$ atoms
139: is divided up into groups of $N$ atoms each. Each such group is then itself
140: further subdivided into $N/d$ equally spaced clumps of $d$ atoms each, with a
141: distance between clumps of $(\mathcal{N}/N)\cdot d$. A family of
142: coarse-grained descriptions is introduced by restricting attention to the
143: average positions of the atoms in a group, which are regarded as the relevant
144: variables defining the system under consideration, and ignoring the internal
145: coordinates within each group, which are regarded as the ``environment''. In
146: the case $d=N$ the $N$ atoms of each group are all neighbours. The
147: corresponding coarse-grained description is therefore entirely local. As $d$
148: decreases from $N$ to $1$ the coarse-grained description becomes more and more
149: nonlocal. In the case $d=1$ the $N$ atoms of each group are dispersed over
150: the whole chain. Brun and Hartle analyse how decoherence, noise and
151: computational complexity of the coarse-grained evolution depends on the
152: nonlocality parameter $d$ and thus show that local coarse-grainings are
153: characterised by a higher degree of classical predictability.
154:
155: The dynamical system studied by Brun and Hartle is linear. In this paper we
156: analyse classical predictability for a family of coarse grainings for a {\em
157: nonlinear\/} chaotic map, the quantum baker's
158: map~\cite{Balazs1989,Saraceno1990}. To quantify predictability, we compute
159: the entropy increase for the evolution: the greater the rate of entropy
160: increase, the less predictable is the evolution. We consider a family of
161: hierarchical multi-scale coarse grainings and show that predictability
162: decreases as the number of scales characterising the coarse-graining increases.
163:
164: The paper is organised as follows. We start with a short introduction
165: to the quantum baker's map (Sec.~\ref{sec:baker}) and the decoherent histories
166: formalism (Sec.~\ref{sec:DecHist}). We then introduce the family of coarse grainings
167: (Sec.~\ref{sec:DifferentCoarse-grainedDescriptions}), describe how the rate
168: of entropy increase depends on the coarse-graining (Sec.~\ref{sec:Results}),
169: and finally present detailed derivations of our results
170: (Sec.~\ref{sec:Derivations}).
171:
172: \section{Background}
173: \subsection{Quantum baker's map} \label{sec:baker}
174:
175: The quantum baker's map \cite{Balazs1989,Saraceno1990} is a prototypical
176: quantum map invented for the theoretical investigation of quantum chaos.
177: It was introduced as a quantised version of the classical
178: baker's transformation \cite{Arnold1968}. There is, however, no unique
179: quantisation procedure \cite{Berry1979}. The original definition of the
180: map~\cite{Balazs1989,Saraceno1990} is based on Weyl's
181: quantisation~\cite{Weyl1950} of the unit square. In
182: \cite{Schack2000a} a class of quantum baker's maps has been defined by
183: exploiting formal similarities between the symbolic dynamics
184: \cite{Alekseev1981} for the classical baker's map on the one hand and the
185: dynamics of strings of quantum bits (qubits) on the other hand. These maps
186: admit a symbolic description in terms of shifts on strings of qubits similar
187: to classical symbolic dynamics~\cite{Alekseev1981}. Their symbolic description
188: has been further developed in~\cite{Soklakov2000a}.
189:
190: Let us give a short introduction following~\cite{Schack2000a}.
191: Quantum baker's maps are defined on the
192: $D$-dimensional Hilbert space of the quantised unit square \cite{Weyl1950}.
193: For consistency of units, we let the quantum scale on ``phase space'' be
194: $2\pi\hbar=1/D$. Following Ref.~\cite{Saraceno1990}, we choose half-integer
195: eigenvalues $q_j=(j+{1\over2})/D$, $j=0,\ldots,D-1$, and $p_k=(k+{1\over2})/D$,
196: $k=0,\ldots,D-1$, of the discrete ``position'' and ``momentum''
197: operators $\hat q$ and $\hat p$, respectively, corresponding to
198: antiperiodic boundary conditions. We further assume that $D=2^N$,
199: which is the dimension of the Hilbert space of $N$ qubits.
200:
201: The $D=2^N$ dimensional Hilbert space modelling the unit square can be
202: identified with the product space of $N$ qubits via
203: \begin{equation}
204: \Ket{q_j} =
205: \Ket{\xi_1}\otimes\Ket{\xi_2}\otimes\cdots\otimes\Ket{\xi_N} \;,
206: \label{eq:tensor1}
207: \end{equation}
208: where $j=\sum_{l=1}^N \xi_l2^{N-l}$, $\xi_l\in\{0,1\}$,
209: and where each qubit has basis states $|0\rangle$ and $|1\rangle$.
210: We can write $q_j$ as a binary fraction, $q_j=0.\xi_1\xi_2\ldots\xi_N1$.
211: Let us define the notation
212: \begin{equation}
213: \Ket{.\xi_1\xi_2\ldots \xi_N} = e^{i\pi/2} \Ket{q_j} \;;
214: \label{eqtensor}
215: \end{equation}
216: see Ref.~\cite{Schack2000a} for the reason for the phase factor $e^{i\pi/2}$.
217: Momentum and position eigenstates are related through the quantum Fourier
218: transform operator $\hat F$ {\cite{Saraceno1990}}, i.e.,
219: $\hat F\Ket{q_k}=\Ket{p_k}$.
220:
221: By applying the Fourier transform operator to the $n$ rightmost bits of
222: the position eigenstate $|.\xi_{n+1}\ldots\xi_N\xi_n\ldots\xi_1\rangle$,
223: one obtains the family of states \cite{Schack2000a}
224: \begin{eqnarray} \label{baker8}
225: |\xi_1\ldots\xi_n.\xi_{n+1}\ldots\xi_N\rangle&\equiv&
226: 2^{-n/2}e^{i\pi(0.\xi_n\ldots\xi_11)}
227: |\xi_{n+1}\rangle
228: \otimes\cdots\otimes
229: |\xi_N\rangle \otimes\cr
230: & & (|0\rangle+e^{2\pi i(0.\xi_11)}|1\rangle)
231: \otimes(|0\rangle+e^{2\pi i(0.\xi_2\xi_11)}|1\rangle)
232: \otimes\cr
233: & & (|0\rangle+e^{2\pi i(0.\xi_3\xi_2\xi_11)}|1\rangle)
234: \otimes\cdots\otimes\cr
235: & & (|0\rangle+e^{2\pi i(0.\xi_n\ldots\xi_11)}|1\rangle)\;,
236: \end{eqnarray}
237: where $1\leq n\leq N-1$.
238: For fixed values of $n$ and $N$ we will use the notation
239: \begin{equation} \label{Eq:basis_n}
240: |\xi_1\ldots\xi_N\rangle_n\equiv
241: |\xi_1\ldots\xi_n.\xi_{n+1}\ldots\xi_N\rangle\,.
242: \end{equation}
243: These states form an orthonormal basis of the Hilbert space.
244: The state (\ref{baker8}) is localised in both
245: position and momentum: it is strictly localised within a position region
246: of width $1/2^{N-n}$, centred at position
247: $q=0.\xi_{n+1}\ldots\xi_N1$, and it is approximately localised within
248: a momentum region of width $1/2^{n}$, centred at momentum
249: $p=0.\xi_n\ldots\xi_11$.
250:
251: For each fixed $n$, $0\leq n\leq N-1$, the quantum
252: baker's map $B_n$ is defined by
253: \begin{equation} \label{baker9}
254: B_n|\xi_1\ldots\xi_n.\xi_{n+1}\ldots\xi_N\rangle =
255: |\xi_1\ldots\xi_{n+1}.\xi_{n+2}\ldots\xi_N\rangle \;,
256: \end{equation}
257: i.e.
258: \begin{equation} \label{baker10}
259: B_n|\xi_1\ldots \xi_N\rangle_n =|\xi_1\ldots\xi_N\rangle_{n+1}\;.
260: \end{equation}
261: \noindent
262: The action of the map $B_n$ on the basis states
263: (\ref{baker8}) is thus given by a shift of the dot by one
264: position. In phase-space language, the map $\hat B_n$ takes a state
265: localised at
266: $(q,p)=(0.\xi_{n+1}\ldots\xi_N1,0.\xi_n\ldots\xi_11)$ to a state localised at
267: $(q',p')=(0.\xi_{n+2}\ldots\xi_N1,0.\xi_{n+1}\ldots\xi_11)$,
268: while it stretches the
269: state by a factor of two in the $q$ direction and squeezes it by a
270: factor of two in the $p$ direction.
271: For $n=N-1$, the map is the original quantum baker's map
272: as defined in Ref.~\cite{Saraceno1990}.
273:
274:
275: For the sake of clarity, it will be convenient to simplify our
276: notation slightly. Throughout the paper $n$ and $N$ are fixed.
277: So we may omit the index $n$ and denote the quantum baker's map
278: simply by $B$, always keeping in mind that we are dealing
279: with the special baker's map $B_n$ for the given value of $n$.
280:
281: \subsection{Decoherent histories formalism} \label{sec:DecHist}
282:
283: The decoherent histories
284: formalism~\cite{Gell-MannHartle1993,Griffiths1984,Omnes1988,
285: Gell-MannHartle1990,DowkerHalliwell1992} provides a framework for
286: investigating classicality in quantum
287: theory~\cite{DowkerHalliwell1992,Gell-MannHartle1993}.
288: The formalism assigns probabilities to quantum histories,
289: i.e.\ ordered sequences of quantum-mechanical ``propositions''.
290: Mathematically, these propositions are represented by projectors.
291: An exhaustive set of mutually exclusive propositions
292: corresponds to a complete set of mutually orthogonal projectors.
293: In this approach to quantum theory a quantum mechanical system is
294: said to exhibit classical behaviour when the probability distribution
295: over histories is strongly peaked about histories
296: having correlations in time implied by classical deterministic
297: laws~\cite{Gell-MannHartle1993,BrunHartle1999-PRD}. Due to quantum
298: interference one cannot always assign probabilities
299: to a set of histories in a consistent way. For this to be possible,
300: the set of histories must be decoherent. Decoherence of histories
301: is therefore a prerequisite for classical behaviour. In general, only
302: coarse-grained sets of histories are decoherent.
303:
304: For our purpose it will be sufficient to consider a slightly
305: simplified version of the general decoherent histories
306: framework, tailored to a system dynamics
307: induced by a fixed unitary quantum map $U$ and restricted to
308: the special but natural case, in which histories are constructed
309: from a fixed exhaustive set of mutually exclusive propositions.
310:
311: A {\em projective partition\/} of a Hilbert space $\cal H$ is
312: a complete set of mutually orthogonal projection operators $\,\{P_{\mu}\}$ on
313: ${\cal H}$, i.e., $P_{\mu}P_{\mu'}=\delta_{\mu\mu'}P_{\mu}\:$ and
314: $\:\sum_{\mu}P_{\mu}=\one_{\cal H}\:$, where $\one_{\cal H}$ denotes
315: the unit operator on $\cal H$.
316: A projective partition is {\em
317: fine-grained\/} if all projectors are one-dimensional,
318: i.e., $\,\forall\,\mu\;$
319: $\mbox{rank}(P_{\mu})=\mbox{dim}\big(\mbox{supp}(P_{\mu})\big)=1$
320: \footnote{The support of a Hermitian operator $A$ is defined to
321: be the vector space spanned by the eigenvectors of $A$ corresponding
322: to its non-zero eigenvalues.}, and {\em coarse-grained} otherwise.
323:
324: Given a projective partition $\{P_{\mu}\}$ of a Hilbert space $\cal H$,
325: a string of length $k$ of projectors $P_{\alpha}\in\{P_{\mu}\}$
326: defines a history of length $k$:
327: \begin{equation}
328: h_{\balphas} \equiv \left(P_{\alpha_1},P_{\alpha_2},\dots,P_{\alpha_k}\right)\;,
329: \end{equation}
330: where $\balpha\equiv\alpha_1\alpha_2\dots\alpha_k$. The set of all
331: such histories, $\mathbb{H}[\{P_{\mu}\}\,;\,k\,]\equiv\big\{h_{\balphas}\,:\:
332: h_{\balphas}\in\{P_{\mu}\}^k\big\}$, forms the
333: exhaustive set of mutually exclusive histories of length
334: $k$. Histories are ordered sequences of projection
335: operators, corresponding to quantum-mechanical propositions.
336: Note that we restrict attention to histories constructed from a
337: fixed exhaustive set of mutually exclusive propositions:
338: the projectors $P_{\alpha_{j}}$ within the
339: sequences are all chosen from the same projective partition,
340: for all times $j=1,\ldots,k$.
341:
342: A set of histories $\mathbb{H}[\{P_{\mu}\}\,;\,k\,]$
343: is called {\em fine-grained\/} ({\em coarse-grained\/}) if it is
344: constructed from a fine-grained (coarse-grained) projective partition.
345: A single history $h_{\balphas}\in\mathbb{H}[\{P_{\mu}\}\,;\,k\,]$
346: is called {\em fine-grained}, if it is represented by
347: a sequence of 1-dimensional projectors, and {\em coarse-grained\/} otherwise.
348:
349: An initial state represented by a density operator $\rho_0$ on ${\cal H}$ and
350: a unitary dynamics generated by a unitary map $U:\cal H\rightarrow\cal H$
351: induce a probabilistic structure on the event algebra associated with
352: $\mathbb{H}[\{P_{\mu}\}\,;\,k\,]$, if the following decoherence conditions are
353: satisfied. These are given in terms of properties of the {\em decoherence
354: functional\/} $\mathcal{D}_{U,\,\rho_0}\,[\cdot,\cdot]$ on
355: $\mathbb{H}[\{P_{\mu}\}\,;\,k\,]\times\mathbb{H}[\{P_{\mu}\}\,;\,k\,]$,
356: defined by
357: \begin{equation}
358: \mathcal{D}_{U,\,\rho_0}\,[h_{\balphas},h_{\bbetas}]\equiv
359: \mbox{Tr}\left[C_{\balphas}\,\rho\,C_{\bbetas}^{\dagger}\right]\:,
360: \end{equation}
361: where
362: \begin{eqnarray} \label{Classoperators}
363: C_{\balphas}\equiv C_{h_{\mbox{\tiny{$\balpha$}}}}
364: & \equiv &\left(U^{\dagger\,k}P_{\alpha_k}U^k\right)
365: \left(U^{\dagger\,k-1}P_{\alpha_{k-1}}U^{k-1}\right)\dots
366: \left(U^{\dagger}P_{\alpha_1}U\right)\nonumber\\
367: &=& U^{\dagger\,k}P_{\alpha_k}UP_{\alpha_{k-1}}U\dots
368: P_{\alpha_2}UP_{\alpha_1}U\;.
369: \end{eqnarray}
370: The set of histories $\mathbb{H}[\{P_{\mu}\}\,;\,k\,]$ is said to be
371: decoherent with respect to a given
372: unitary map $U:\cal H\rightarrow\cal H$ and a given initial
373: state $\rho_0$, if
374: \begin{equation} \label{eq:consistency}
375: \mathcal{D}_{U,\,\rho_0}\,[h_{\balphas},h_{\bbetas}]
376: \propto \delta_{\balphas\bbetas}\equiv
377: \prod_{j=1}^k\delta_{\alpha_j \beta_j}
378: \end{equation}
379: for all $h_{\balphas},h_{\bbetas}\in
380: \mathbb{H}[\{P_{\mu}\}\,;\,k\,]$.
381: If this decoherence condition is satisfied, the diagonal elements
382: of the decoherence functional,
383: $p[h_{\balphas}]=\mathcal{D}_{U,\,\rho_0}\,[h_{\balphas},h_{\balphas}]$,
384: can be interpreted as the probabilities of the histories.
385: For a decoherent set of histories, the entropy, $H[\{h_{\balphas}\}]$,
386: can be defined as follows \cite{Gell-MannHartle1990,Hartle1998,BrunHartle1999-PRE}:
387: \begin{eqnarray} \label{def:Entropy}
388: H[\{h_{\balphas} \}]
389: &\equiv& -\sum_{\balphas}p[h_{\balphas}]\log_2p[h_{\balphas}]\nonumber\\
390: &=&-\sum_{\balphas}\mathcal{D}_{U,\,\rho_0}\,[h_{\balphas},h_{\balphas}]
391: \log_2\Big(\mathcal{D}_{U,\,\rho_0}\,[h_{\balphas},h_{\balphas}]\Big)\;.
392: \end{eqnarray}
393:
394:
395: \section{Predictability for different coarse-grainings of the quantum baker's map }
396: \label{sec:coarse}
397:
398: This section is organised as follows.
399: Subsection~\ref{sec:DifferentCoarse-grainedDescriptions}, which discusses
400: coarse-grained descriptions of the quantum baker's map, contains two parts:
401: part~1 introduces a family of coarse-grained projective partitions of the
402: Hilbert space, which are then used in part~2 to construct a class of
403: coarse-grained sets of histories. Subsection~\ref{sec:Results}
404: summarises the main results of this paper, which are then derived and
405: illustrated in detail in Subsection~\ref{sec:Derivations}.
406:
407: \vspace{-2mm}
408: \subsection{Coarse-grainings}
409: \label{sec:DifferentCoarse-grainedDescriptions}
410: \vspace{-1mm}
411:
412: \subsubsection{Coarse-grained partitions}
413: \label{sec:Coarse-grained_partitions}
414:
415: Let us first introduce two different types of coarse-grained
416: projective partitions of the $2^N$-dimensional Hilbert space modelling
417: the unit square, which later will be regarded as special cases
418: of a family of more general coarse-grained descriptions.
419: We refer to the definitions and notations of Sec.~\ref{sec:baker}.
420: In particular we use the orthonormal basis (\ref{Eq:basis_n}) of
421: the Hilbert space to construct the partitions.
422:
423: For a fixed binary string $\y=y_1\ldots y_{N-l-r}\in \{0,1\}^{N-l-r}$
424: we define the ``local'' projection operators by
425: \begin{equation} \label{coarseProjectorsloc}
426: {P}_{\ys}^{(l,r)}\equiv
427: \sum_{a_1,\ldots,a_l\atop b_1,\ldots,b_r}
428: |a_1\dots a_l\ \y\ b_1\dots b_r\rangle_n\,_n
429: \langle a_1\dots a_l\ \y\ b_1\dots b_r|\,
430: \equiv\sum_{\as\in \{0,1\}^{l} \atop \bs\in \{0,1\}^{r}}
431: |\at\ \y\ \bt\ \rangle_n\,_n
432: \langle \at\ \y\ \bt|\,,
433: \end{equation}
434: \noindent
435: and for fixed strings $\y^1\in \{0,1\}^{s_1}$ and $\y^2\in
436: \{0,1\}^{s_2}$ we define the ``nonlocal''
437: projection operators by
438: \begin{eqnarray} \label{CheckerboardProjectors}
439: {P}_{\ys^1,\,\ys^2}^{(l,m_l, m_r, r)}&\equiv &
440: \sum_{\as\in \{0,1\}^{l}}\:\sum_{\bs\in \{0,1\}^{r}}\:
441: \sum_{\,\xibolds\in \{0,1\}^{m_l+m_r}}|\at\ \y^1\ \xibold\ \y^2\ \bt\ \rangle_n\,_n
442: \langle\at\ \y^1\ \xibold\ \y^2\ \bt\ |\;\\
443: &\equiv&\sum_{a_1,\ldots,a_l\atop b_1,\ldots,b_r}
444: \sum_{\xi_1\ldots\xi_{m_l}\atop \xi_{m_l+1}\ldots\xi_{m_l+m_r}}
445: |a_1\dots a_l\ \y^1\ \xi_1\ldots\xi_{m_l}\,.\,\xi_{m_l+1}\ldots\xi_{m_l+m_r}
446: \y^2\ b_1\dots b_r\rangle\times\nonumber\\ && \hspace{3.5cm}\langle a_1\dots a_l\ \y^1\
447: \xi_1\ldots\xi_{m_l}\,.\,\xi_{m_l+1}\ldots\xi_{m_l+m_r} \y^2\ b_1\dots
448: b_r|\nonumber
449: \end{eqnarray}
450: What the terms ``local'' and ``non-local'' mean
451: in this context, will be explained below.
452: Throughout this paper, bold variables denote binary strings.
453: Furthermore, lower indices label individual bits of a
454: string, whereas upper indices will label different strings.
455: It will be convenient to abbreviate a substring
456: $\alpha_{\kappa}\dots\alpha_{\sigma}$ of a string
457: $\balpha=\alpha_1\dots\alpha_{\kappa}\alpha_{\kappa+1}\dots\alpha_{\sigma}
458: \alpha_{\sigma+1}\dots\alpha_{\gamma}$ by $\balpha_{\kappa:\sigma}$.
459: Concatenation of strings is defined in the usual way.
460: Taking the just mentioned example we can, for instance, express
461: the string $\balpha$ as a concatenation of three substrings,
462: $\balpha=\balpha_{1:\kappa-1}\balpha_{\kappa:\sigma}\balpha_{\sigma+1:\gamma}$.
463: The length of a string $\balpha$ will be denoted by $|\balpha|$.
464:
465: For simplicity, we will always assume in the following
466: that $l<n$ and $r<N-n$ in the first case, and
467: $l+s_1\le n$ and $r+s_2\le N-n$ in the second case.
468: In both cases $l$ and $r$ acquire the specific
469: meaning as the number of ``momentum'' and ``position'' bits
470: ignored in the coarse-graining. In the second case,
471: in addition $m_l$ most significant momentum
472: bits and $m_r$ most significant position
473: bits are coarse-grained over.
474:
475: The operator ${P}_{\ys}^{(l,r)}$ is a projector on a
476: $2^{l+r}$-dimensional subspace labelled by the string $\y$.
477: The projector ${P}_{\ys^1,\,\ys^2}^{(l,m_l, m_r, r)}$ projects
478: on a $2^{l+m_l+m_r+r}$-dimensional subspace labelled by the pair
479: of strings $(\y^1, \y^2)$.
480: In both cases we are dealing with complete sets of
481: mutually orthogonal projectors, i.e., with projective partitions, as
482: \begin{equation}
483: {P}_{\ys}^{(l,r)}{P}_{\ys'}^{(l,r)}=\delta_{\ys,\ys'}{P}_{\ys}^{(l,r)}\quad
484: \mbox{and}\quad\sum_{\ys}{P}_{\ys}^{(l,r)}=\one\;,
485: \end{equation}
486: \begin{equation}
487: {P}_{\ys^1,\,\ys^2}^{(l,m_l, m_r, r)}{P}_{\ys^{'1},\,\ys^{'2}}^{(l,m_l, m_r, r)}=
488: \delta_{\ys^1,\,\ys^{'1}}\delta_{\ys^2,\,\ys^{'2}}
489: {P}_{\ys^1,\,\ys^2}^{(l,m_l, m_r,r)}\quad\mbox{and}\quad
490: \sum_{\ys^1,\,\ys^2}{P}_{\ys^1,\,\ys^2}^{(l,m_l, m_r, r)}=\one\;.
491: \end{equation}
492:
493:
494: Let us explain what is meant by ``local'' and ``nonlocal''
495: regarding the just introduced projection operators.
496: The projection operators ${P}_{\ys}^{(l,r)}$ and
497: ${P}_{\ys^1,\,\ys^2}^{(l,m_l, m_r, r)}$
498: project on subspaces of the Hilbert space associated
499: with phase-space regions of the unit square in which
500: the corresponding eigenstates with eigenvalue 1 are localised.
501: In the case of the projectors ${P}_{\ys}^{(l,r)}$
502: these regions are connected cells whose location
503: within the unit square of the phase space is determined
504: by the specified most significant position and momentum
505: bits given by the binary string $\y=y_1\ldots y_{N-l-r}$.
506: The size of these cells depends on the significance of
507: the scales which are not resolved and therefore ignored, i.e.\ coarse-grained over.
508: In the case of the projectors ${P}_{\ys^1,\,\ys^2}^{(l,m_l, m_r, r)}$,
509: on the other hand, there is coarse-graining also at
510: the most significant scales: a number of the most significant
511: position and momentum bits are not specified. The associated phase space domains
512: must therefore consist of disconnected parts spread over the whole
513: unit square, the number depending on how many most significant
514: position and momentum bits are coarse-grained over, i.e.\ on the
515: parameter $m\equiv m_l+m_r$. For an illustration see Fig.~\ref{fig1}.
516:
517: \begin{figure}[bt]
518: \begin{center}
519: \includegraphics[width=10cm]{checkerb.eps}
520: \bigskip
521: \bigskip
522: \caption{A schematic illustration of the projectors ${P}_{\ys}^{(l,r)}$ and
523: ${P}_{\ys^1,\,\ys^2}^{(l,m_l, m_r, r)}$. {\bf (a)} Let
524: $n-l=2$, $N-l-r=4$, and $\y=y_1\ldots y_4=1110$. The projector
525: ${P}_{\ys=1110}^{(l,r)}$ is then in approximate correspondence with
526: the phase space region shaded in grey. {\bf (b)} Let $n-l=4$, $m_l=2$,
527: $m_r=2$, $\y^1=10$ and $\y^2=01$. The projector
528: ${P}_{\ys^1=10,\,\ys^2=01}^{(l,2,2, r)}$ is then in approximate
529: correspondence with the disconnected phase space region given by the 16
530: black cells.}
531: \label{fig1}
532: \end{center}
533: \end{figure}
534:
535:
536: We will also use the diagram notation for the introduced projectors:
537: \begin{eqnarray}
538: \label{coarseProjectorsloc_diagram}
539: {P}_{\ys}^{(l,r)}&\equiv&(\,\underbrace{\Box\Box\dots\Box}_{l}
540: \ \y\ \underbrace{\Box\Box\dots \Box}_{r}\,)\;,\\
541: \label{CheckerboardProjectors_diagram}
542: {P}_{\ys^1,\,\ys^2}^{(l,m_l, m_r, r)}&\equiv& (\,\underbrace{\Box\Box\dots\Box}_{l}
543: \ \y^1\ \underbrace{\Box\Box\dots\Box}_{m_l}\ .\ \underbrace{\Box\Box\dots \Box}_{m_r}\
544: \y^2\ \underbrace{\Box\Box\dots \Box}_{r}\,)\;,
545: \end{eqnarray}
546: where the empty boxes indicate the bits which are coarse-grained
547: over. We can write the projectors of the second type as
548: sums over projectors of the first type:
549: \begin{equation} \label{nonlocal=sumoverlocal}
550: {P}_{\ys^1,\,\ys^2}^{(l,m_l, m_r, r)}= \sum_{\,\xibolds\in \{0,1\}^{m_l+m_r}}
551: {P}_{\ys^1\xibolds\ys^2}^{(l,r)}\;,
552: \end{equation}
553: where $\y^1\xibold\y^2$ means the concatenation of the three strings
554: $\y^1$, $\xibold$ and $\y^2$. Remember that in the definition of
555: the projectors ${P}_{\ys^1,\,\ys^2}^{(l,m_l, m_r, r)}$
556: we assume that $l+|\y^1|\le n$ and
557: $r+|\y^2|\le N-n$.
558:
559: The projectors (\ref{coarseProjectorsloc_diagram}) and
560: (\ref{CheckerboardProjectors_diagram}) are special cases
561: of the family of all projection operators, which define the
562: scales at which information is lost in the symbolic
563: representation. In general such projectors exhibit
564: structure on many different scales, and the most
565: general projector of this type would be of the form
566: \begin{equation}
567: \label{multi-scales_coarse-graining_diagram}
568: {P}_{\ys^1,\,\ys^2,\dots,\ys^{\lambda}}^{(l,m_1, m_2,\dots,m_{\lambda-1},r)}
569: =
570: \bigg( \,\underbrace{\Box\dots\Box}_{l}
571: \ \y^{1}\ \underbrace{\Box\dots\Box}_{m_1}\,\y^{2}\
572: \underbrace{\Box\dots\Box}_{m_2}\ \,\dots\dots\,
573: \y^{\lambda-1}\ \underbrace{\Box\dots\Box}_{m_{\lambda-1}}\
574: \y^{\lambda}\ \underbrace{\Box\dots\Box}_{r}\,\bigg)\;.
575: \end{equation}
576: The projector (\ref{multi-scales_coarse-graining_diagram})
577: defines a coarse-graining in which information is lost on
578: several different scales. We will call this a {\em multi-scale
579: coarse-graining\/} or {\em hierarchical coarse-graining}.
580: Accordingly, the special cases
581: (\ref{coarseProjectorsloc_diagram}) and
582: (\ref{CheckerboardProjectors_diagram}) will be called
583: 1-scale and 2-scale coarse-graining, respectively.
584: The 2-scale coarse-graining (\ref{CheckerboardProjectors_diagram})
585: we introduced above is a special 2-scale coarse-graining,
586: as we assumed that the coarse-grained {\em island\/} of
587: size $m_l+m_r$ between the specified strings $\y^1$ and
588: $\y^2$ lies around the dot separating the momentum and
589: position bits in the symbolic representation. The first step
590: towards a generalisation is to combine the two parameters
591: $m_l$ and $m_r$ (i.e.\ the number of
592: most significant momentum and position bits that
593: are coarse-grained over in the symbolic representation)
594: to a single parameter $m=m_l+m_r$ and allow the corresponding
595: coarse-grained island of size $m$ between the specified strings $\y^1$ and
596: $\y^2$ to lie anywhere, not necessarily at
597: the most significant region around the dot.
598: The next step is to introduce several
599: coarse-grained islands of this kind, on several scales.
600: An event will then be specified by bit strings $\y^1$, $\y^2$,
601: $\dots$, $\y^{\lambda}$ of length $|\y^i|=s_i$ at a time, separated by
602: $(\lambda-1)$ coarse-grained islands of size $m_i$ each, where $\lambda>1$,
603: as in Eq.~(\ref{multi-scales_coarse-graining_diagram}).
604:
605: More precisely, the most general family of coarse-grained descriptions is
606: represented by sets of projection operators defined as follows:
607: \begin{eqnarray} \label{multi-scales_coarse-graining_projectors}
608: {P}_{\ys^1,\,\ys^2,\dots,\ys^{\lambda}}^{(l,m_1, m_2,\dots,m_{\lambda-1},r)}&\equiv &
609: \sum_{\as\in \{0,1\}^{l}}\:\sum_{\bs\in \{0,1\}^{r}}\:
610: \sum_{\,\xibolds^1\in \{0,1\}^{m_1}}\dots
611: \sum_{\,\xibolds^{\lambda-1}\in \{0,1\}^{m_{\lambda-1}}}\nonumber\\
612: &&|\at\ \y^1\ \xibold^1\ \y^2\ \xibold^2\ \dots \xibold^{\lambda-1}
613: \y^{\lambda}\bt\ \rangle_n\,_n\langle
614: \at\ \y^1\ \xibold^1\ \y^2\ \xibold^2\ \dots
615: \xibold^{\lambda-1}\y^{\lambda}\bt\ |\nonumber\\
616: &=&
617: \sum_{\,\xibolds^1\in \{0,1\}^{m_1}}\dots
618: \sum_{\,\xibolds^{\lambda-1}\in \{0,1\}^{m_{\lambda-1}}}
619: {P}_{\ys^1\xibolds^1\ys^2\xibolds^2\dots
620: \ys^{\lambda-1}\xibolds^{\lambda-1}\ys^{\lambda}}^{(l,r)}\;,
621: \end{eqnarray}
622: \noindent
623: where $\y^1\xibold^1\y^2\xibold^2\dots
624: \y^{\lambda-1}\xibold^{\lambda-1}\y^{\lambda}$
625: means the concatenation of the particular strings
626: $\y^1$, $\xibold^1$, $\y^2,\dots,\xibold^{\lambda-1}$,
627: $\y^{\lambda}$. We still assume $l<n$ and $r<N-n\,$.
628: Eq.~(\ref{multi-scales_coarse-graining_diagram})
629: is a diagram notation of
630: Eq.~(\ref{multi-scales_coarse-graining_projectors}).
631: It is easily seen that for fixed $m_1,\dots,m_{\lambda-1}$ the
632: set $\;\{{P}_{\ys^1,\,\ys^2\,,\dots ,\, \ys^{\lambda}}^{(l,m_1,
633: m_2\,,\dots ,\,m_{\lambda-1}, r)}\}$ forms
634: a projective partition of the Hilbert space, as
635: \begin{eqnarray}
636: {P}_{\ys^1,\,\ys^2\,,\dots ,\, \ys^{\lambda}}^{(l,m_1,
637: m_2\,,\dots ,\,m_{\lambda-1}, r)}
638: {P}_{\ys^{'1},\,\ys^{'2}\,,\dots ,\, \ys^{'\lambda}}^{(l,m_1,
639: m_2\,,\dots ,\,m_{\lambda-1}, r)}&=&
640: \delta_{\ys^1,\,\ys^{'1}}\delta_{\ys^2,\,\ys^{'2}}\times\dots
641: \times\delta_{\ys^{\lambda},\,\ys^{'\lambda}}
642: {P}_{\ys^1,\,\ys^2\,,\dots ,\,\ys^{\lambda}}^{(l,m_1,
643: m_2\,,\dots ,\,m_{\lambda-1}, r)}\nonumber\\
644: \quad\mbox{and}\quad
645: \sum_{\ys^1,\,\ys^2,\dots, \ys^{\lambda}}
646: {P}_{\ys^1,\,\ys^2\,,\dots ,\, \ys^{\lambda}}^{(l,m_1,
647: m_2\,,\dots ,\,m_{\lambda-1}, r)}&=&\one\;.
648: \end{eqnarray}
649:
650:
651: \subsubsection{Coarse-grained histories}
652:
653: In order to investigate coarse-grained evolution we now construct
654: coarse-grained histories. By considering different types of histories
655: constructed from different types of coarse-grained projective partitions
656: we obtain different coarse-grained effective evolutions.
657: Our investigation of the coarse-grained evolution of the
658: quantum baker's map starts with
659: the special cases of 1-scale and 2-scale coarse-grainings as
660: defined in Eqs.~(\ref{coarseProjectorsloc_diagram}) and
661: (\ref{CheckerboardProjectors_diagram}). We first compare
662: the different members of the family
663: \begin{eqnarray}
664: \label{family_ofcoarse-grainings:set}
665: \bigg\{\left\{{P}_{\ys^1,\,\ys^2}^{(l,m_l, m_r, r)}\,:\,
666: \y^1\in \{0,1\}^{s_1}\,,\,\y^2\in
667: \{0,1\}^{s_2}\right\}\,&:&\:l,r,m_l,m_r,s_1,s_2\in
668: \{0,1,2,\dots\}\;\;\phantom{\bigg\}}\nonumber\\
669: \phantom{\bigg\{}\mbox{such that}\quad l+s_1\le n\,,\,r+s_2\le N-n
670: &\mbox{and}&
671: l+r+s_1+s_2+ m_l+m_r=N\bigg\}\nonumber\\&&
672: \end{eqnarray}
673: of coarse-grained descriptions, parameterised by
674: $l$, $r$, $s_1$, $s_2$, $m_l$ and $m_r$,
675: with respect to predictability of the evolution.
676: Our results will concern only such members of this family
677: for which $s_1$ and $s_2$ are significantly greater than $1$, and $s_1\ge m_l+m_r$.
678: Furthermore, in order to obtain the classical limit of the quantum
679: baker's map, we will be considering only members with very large
680: value for the parameter $l$, as $\hbar\rightarrow 0$ will correspond
681: to $l \rightarrow \infty$. Finally, the results will show that only
682: $m=m_l+m_r$ matters, and the specification ``$m_l$ most
683: significant momentum bits and $m_r$ most significant position bits are
684: coarse-grained over'' therefore be unnecessary. Note that the local 1-scale
685: coarse-graining (\ref{coarseProjectorsloc}) is included in
686: this family as the special case $m_l+m_r=0$.
687:
688: The histories corresponding to 1-scale and 2-scale coarse-graining
689: (\ref{coarseProjectorsloc_diagram}) and
690: (\ref{CheckerboardProjectors_diagram}) will be
691: labelled by finite sequences of strings in the first case and pairs of
692: finite sequences of strings in the second case, respectively:
693: \begin{eqnarray} \label{LocalHistories}
694: h_{\vec{\ys}}&\equiv&
695: \big({P}^{(l,r)}_{\ys^1},{P}^{(l,r)}_{\ys^2},
696: \dots,{P}^{(l,r)}_{\ys^k}\big)_{\phantom{|_{|_{|_{|_|}}}}}\;,
697: \end{eqnarray}
698: where $\vec{\y}= (\y^1, \dots ,\y^k)$ is a sequence of strings
699: $\y^j\in\{0,1\}^{N-l-r}$, $j=1, \dots , k$;
700: \begin{eqnarray} \label{CheckerboardHistories}
701: h_{\vec{\ys}^1,\,\vec{\ys}^2}&\equiv&
702: \big({P}_{\ys^{1,1},\,\ys^{1,2}}^{(l,m_l, m_r, r)},
703: {P}_{\ys^{2,1},\,\ys^{2,2}}^{(l,m_l, m_r, r)}, \dots,
704: {P}_{\ys^{k,1},\,\ys^{k,2}}^{(l,m_l, m_r,
705: r)}\big)_{\phantom{|_{|_{|_{|_|}}}}}\;,
706: \end{eqnarray}
707: where $(\vec{\y}^1,\vec{\y}^2) = ((\y^{1,1}, \dots ,\y^{k,1}),(\y^{1,2}, \dots ,\y^{k,2}))$
708: is a pair of finite sequences of strings $\y^{j,i}\in\{0,1\}^{s_i}$,
709: $j=1, \dots , k$, $i=1,2$, labelling the history.
710:
711:
712: To examine decoherence of this set of histories and calculate its probability
713: distribution we will evaluate the decoherence functional
714: \begin{equation} \label{dfunc_local}
715: {\mathcal D}_{B,\,\rho_0}[h_{\vec{\ys}},
716: h_{\vec{\zs}} ]
717: =
718: \Tr[P^{(l,r)}_{\ys^k}B P^{(l,r)}_{\ys^{k-1}}
719: B\cdots P^{(l,r)}_{\ys^1}
720: B\rho_0 B^\dag P^{(l,r)}_{\zs^1}
721: \cdots
722: B^\dag P^{(l,r)}_{\zs^{k-1}}B^\dag P^{(l,r)}_{\zs^k}] \;,
723: \end{equation}
724: and
725: \begin{equation*} \label{checkerboard-dfunc}
726: {\mathcal D}_{B,\,\rho_0}[h_{\vec{\ys}^1,\,\vec{\ys}^2},
727: h_{\vec{\zs}^1,\,\vec{\zs}^2} ]
728: \equiv \hspace{11cm}
729: \end{equation*}
730: \begin{equation} \label{checkerboard-dfunc}
731: = \Tr[{P}_{\ys^{k,1},\,\ys^{k,2}}^{(l,m_l, m_r, r)}B
732: {P}_{\ys^{k-1,1},\,\ys^{k-1,2}}^{(l,m_l, m_r, r)} B\cdots
733: {P}_{\ys^{1,1},\,\ys^{1,2}}^{(l,m_l, m_r, r)}
734: B\rho_0 B^\dag
735: {P}_{\zs^{1,1},\,\zs^{1,2}}^{(l,m_l,m_r,r)}B^\dag\cdots
736: B^\dag
737: {P}_{\zs^{k,1},\,\zs^{k,2}}^{(l,m_l, m_r, r)}] \;,
738: \end{equation}
739: respectively.
740:
741:
742: Whether the decoherence functional is diagonal or not depends
743: on the initial state $\rho_0$. In order to check decoherence
744: of a given set of histories and assign probabilities
745: to them we therefore need to specify the initial state from which
746: the histories start.
747: Here we choose a certain class of states as the initial states
748: for the histories, namely the discrete set of states that are
749: induced via normalisation by the set of projectors defining the histories.
750: We therefore assume the initial state $\rho_0$ to be of the same form
751: as the events in the histories, i.e.\ to be proportional to one of
752: the projection operators of the set $\{ {P}_{\ys}^{(l,r)}\}$ or
753: $\{ {P}_{\ys^1,\,\ys^2}^{(l,m_l, m_r, r)}\}$, respectively:
754: \begin{equation}
755: \label{LocalInitialstate}
756: \rho_0 = \rho_{\xs}^{(l, r)}\equiv
757: 2^{-(l+r)} {P}_{\xs}^{(l,r)}\\
758: \equiv 2^{-(l+r)} (\,\underbrace{\Box\Box\dots\Box}_{l}
759: \ \x\ \underbrace{\Box\Box\dots \Box}_{r}\,)\;,
760: \end{equation}
761: or
762: \begin{eqnarray} \label{CheckerboardInitialstate}
763: \rho_0 &=& \rho_{\xs^1,\,\xs^2}^{(l,m_l, m_r, r)}\equiv
764: 2^{-(l+m_l+m_r+r)} {P}_{\xs^1,\,\xs^2}^{(l,m_l, m_r, r)}\\
765: &\equiv& 2^{-(l+m_l+m_r+r)} (\,\underbrace{\Box\Box\dots\Box}_{l}
766: \ \x^1\ \underbrace{\Box\Box\dots\Box}_{m_l}\ .\
767: \underbrace{\Box\Box\dots \Box}_{m_r}\
768: \x^2\ \underbrace{\Box\Box\dots \Box}_{r}\,)\;.\nonumber
769: \end{eqnarray}
770: The normalisation factor $2^{-(l+r)}$ or $2^{-(l+m_l+m_r+r)}$,
771: respectively, ensures that
772: $\rho_0$ is a density operator, i.e.\ $\Tr[\rho_0]=1$.
773: All calculations in Sec.~\ref{sec:Derivations} will be based
774: on this choice for the initial states, which we regard
775: as the most natural choice within our framework of sets of
776: histories constructed from a given, fixed projective
777: partition.
778:
779: We now generalise the family of sets of coarse-grained
780: histories from the 1-scale and 2-scale coarse-grained
781: descriptions considered above to the general case
782: of multi-scale (or hierarchical) coarse-grainings. The corresponding
783: projective partitions have already been introduced
784: in Eqs.~(\ref{multi-scales_coarse-graining_diagram})
785: and (\ref{multi-scales_coarse-graining_projectors}).
786: The generalised family of coarse-grained descriptions
787: is therefore given by the set:
788: \begin{eqnarray}
789: \label{family_hierarchical_coarse_grainings}
790: \bigg\{\left\{{P}_{\ys^1,\,\ys^2\,,\dots ,\, \ys^{\lambda}}^{(l,m_1,
791: m_2\,,\dots ,\,m_{\lambda-1}, r)}\,\right\}_{\y^j\in \{0,1\}^{s_j}}\,&:&\:\:
792: l,r,m_j, s_j\in\{0,1,2,\dots\}\phantom{\bigg\}}
793: \,,\;\lambda\in\{1,2,3,\dots\}\nonumber\\
794: \phantom{\bigg\{}\mbox{such that}
795: &&l+r+\sum_{j=1}^{\lambda-1} m_j+\sum_{j=1}^{\lambda}s_j=N
796: \bigg\} \;.
797: \end{eqnarray}
798: The members of this family are represented by coarse-grained projective partitions
799: displaying coarseness on several different scales in the symbolic
800: representation. The family is parameterised by $l,r, m_1,\dots
801: ,\,m_{\lambda-1}$, $s_1,\dots,s_{\lambda}$ and $\lambda$
802: with the constraint
803: $l+r+\sum_{j=1}^{\lambda-1} m_j+\sum_{j=1}^{\lambda}s_j=N$.
804: Again, our results will involve only such members of this family,
805: for which $s_1,\dots,s_{\lambda}$ have values significantly greater
806: than $1$, and the value of $l$ is very large (classical limit).
807:
808: Our generalised type of histories is labelled by (finite) sequences
809: of finite sequences of binary strings:
810: \begin{equation}\label{set:hierarchically-coarse-grained-histories}
811: \bigg\{h_{\vec{\ys}^1,\,\vec{\ys}^2,\dots,\vec{\ys}^{\lambda}}\;:\;\vec{\y}^{\,i}=
812: (\y^{1,i}, \dots ,\y^{k,i})\quad\mbox{with}\quad \y^{j,i}\in\{0,1\}^{s_i}\;,\;
813: j=1,\dots,k\;,\; i=1,\dots,\lambda\bigg\}\;.
814: \end{equation}
815: They are explicitely defined by time-ordered sequences of
816: (\ref{multi-scales_coarse-graining_projectors})-type projection
817: operators:
818: \begin{equation}
819: h_{\vec{\ys}^1,\,\vec{\ys}^2,\dots,\vec{\ys}^{\lambda}}
820: \equiv
821: \Big({P}_{\ys^{1,1},\,\ys^{1,2},\dots,\ys^{1,\lambda}}
822: ^{(l,m_1, m_2,\dots,m_{\lambda-1},r)}\;,\;
823: {P}_{\ys^{2,1},\,\ys^{2,2},\dots,\ys^{2,\lambda}}
824: ^{(l,m_1, m_2,\dots,m_{\lambda-1},r)}\;,\dots,\;
825: {P}_{\ys^{k,1},\,\ys^{k,2},\dots,\ys^{k,\lambda}}
826: ^{(l,m_1,
827: m_2,\dots,m_{\lambda-1},r)}\Big)_{\phantom{|_{|_{|_{|_|}}}}} \;.
828: \end{equation}
829:
830: To examine decoherence of the set of
831: histories~(\ref{set:hierarchically-coarse-grained-histories}) and calculate
832: its probability distribution we will evaluate the decoherence functional
833: \begin{equation*}
834: \hspace*{-10cm}{\mathcal
835: D}_{B,\,\rho_0}[h_{\vec{\ys}^1,\,\vec{\ys}^2,\dots,\vec{\ys}^{\lambda}},
836: h_{\vec{\zs}^1,\,\vec{\zs}^2,\dots,\vec{\zs}^{\lambda}} ]
837: \equiv
838: \end{equation*}
839: \begin{eqnarray}
840: \label{hierarchical-dfunc}
841: &=\mbox{Tr}\Big[{P}_{\ys^{k,1},\,\ys^{k,2},\dots,\ys^{k,\lambda}}
842: ^{(l,m_1,
843: m_2,\dots,m_{\lambda-1},r)}B
844: {P}_{\ys^{k-1,1},\,\ys^{k-1,2},\dots,\ys^{k-1,\lambda}}
845: ^{(l,m_1,
846: m_2,\dots,m_{\lambda-1},r)}B\cdots
847: {P}_{\ys^{1,1},\,\ys^{1,2},\dots,\ys^{1,\lambda}}
848: ^{(l,m_1,
849: m_2,\dots,m_{\lambda-1},r)}B\rho_0 B^\dag\times&\nonumber\\
850: &\times {P}_{\zs^{1,1},\,\zs^{1,2},\dots,\zs^{1,\lambda}}
851: ^{(l,m_1,
852: m_2,\dots,m_{\lambda-1},r)}B^\dag\cdots
853: {P}_{\zs^{k-1,1},\,\zs^{k-1,2},\dots,\zs^{k-1,\lambda}}
854: ^{(l,m_1,
855: m_2,\dots,m_{\lambda-1},r)}B^\dag
856: {P}_{\zs^{k,1},\,\zs^{k,2},\dots,\zs^{k,\lambda}}
857: ^{(l,m_1,
858: m_2,\dots,m_{\lambda-1},r)}\Big] \;,&
859: \end{eqnarray}
860: Again we will choose the initial state to be proportional to one
861: of the projection operators defining our coarse-grained
862: description, i.e.\ to one of the
863: (\ref{multi-scales_coarse-graining_projectors})-type projectors:
864: \begin{equation}
865: \label{Hierarchical_coarse-grainingInitialstate}
866: \rho_0 = \rho_{\xs^1,\,\xs^2,\dots,\xs^{\lambda}}^{(l,m_1, m_2,\dots,m_{\lambda-1},r)}\equiv
867: 2^{-(l+r+m_1+m_2+\dots+m_{\lambda-1})}
868: {P}_{\xs^1,\,\xs^2,\dots,\xs^{\lambda}}^{(l,m_1, m_2,\dots,m_{\lambda-1},r)}\;,
869: \end{equation}
870: with the normalisation factor ensuring $\Tr[\rho_0]=1$.
871:
872:
873:
874: \subsection{Main results}
875: \label{sec:Results}
876:
877: To characterise and quantify predictability, we use the rate of the entropy
878: production. The greater the rate of the entropy production, the more
879: unpredictable is the evolution.
880: We begin by stating the results for the
881: family~(\ref{family_ofcoarse-grainings:set}) of 1-scale and 2-scale
882: coarse-grainings.
883: First of all we find that in the asymptotic limit $l\to \infty$
884: all the corresponding members of this family
885: (i.e., all members with very large parameter value $l$),
886: provided that $m_l+m_r$ is finite,
887: lead to decoherent sets of histories, which is the
888: prerequisite for classicality. For finite, but very large $l$
889: the decoherence functional is approximately diagonal, which
890: means approximate decoherence of histories.
891: For very large $l$, the diagonal elements of the decoherence
892: functional, $\mathcal{D}_{B,\,\rho_0}[h_{\vec{\ys}^1,\,\vec{\ys}^2},
893: h_{\vec{\ys}^1,\,\vec{\ys}^2} ]$, may therefore be interpreted
894: as probabilities of the corresponding histories.
895: Furthermore we find that for very large $l$,
896: for all members of the corresponding subset within this family,
897: for which $s_1$ and $s_2$ are significantly greater than $1$,
898: the probabilities of the individual alternative histories
899: of a set are peaked at histories which display
900: regularities according to the {\em classical shift property}.
901:
902: We have compared the rates of entropy increase of the different sets
903: within the family~(\ref{family_ofcoarse-grainings:set}) of
904: coarse-grainings. The result for the local coarse-graining
905: (\ref{coarseProjectorsloc}), i.e.\ for the case $m_l+m_r=0$,
906: was obtained in an earlier work of two of
907: us~\cite{Soklakov2002}. In~\cite{Soklakov2002} it was
908: shown that in this case the coarse-grained quantum baker's map
909: exhibits a linear entropy increase at an asymptotic rate given by the
910: Kolmogorov-Sinai entropy~\cite{Alekseev1981} of the classical
911: chaotic baker's map, namely 1 bit per iteration step:
912: \begin{equation} \label{EntropyLocal}
913: H[\{h_{\vec{\ys}} \}] = k+
914: O(\frac{(l+r-k)\log_2(l+r-k)}{2^{l-2(k^2+k)}})\,,
915: \end{equation}
916: where the set $\{h_{\vec{\ys}} \}$ consists of histories of
917: length $k$.
918:
919: For nonlocal coarse-grainings $m_l+m_r\not=0$,
920: the derivation in the next section give these results:
921: \begin{itemize}
922: \item Entropy after $k$ iteration steps in case $k\le m_l+m_r$:
923: \begin{equation}
924: H[\{ h_{\vec{\ys}^1,\,\vec{\ys}^2}\}] =
925: 2k\;+\;{\cal
926: O}(\frac{(l+r-k)\log_2(l+r-k)}{2^{l-2(k^2+(1+m_l+m_r)k)}})\;
927: \end{equation}
928: \item Entropy after $k$ iteration steps in case $k\ge m_l+m_r$:
929: \begin{equation}
930: H[\{ h_{\vec{\ys}^1,\,\vec{\ys}^2}\}]
931: =k+\;(m_l+m_r)\;+\;{\cal
932: O}(\frac{(l+r-k)\log_2(l+r-k)}{2^{l-2(k^2+(1+m_l+m_r)k)}})\;.
933: \end{equation}
934: \end{itemize}
935:
936: The entropy increase is 2 bits per iteration step as long as the number of
937: iterations $k$ of the quantum baker's map is smaller than $m=m_l+m_r$. As
938: soon as the number of iterations exceeds the parameter $m$, the rate of
939: entropy increase drops to 1 bit per iteration step. Both short-term and
940: long-term rates of entropy increase are thus independent of the
941: non-locality parameter $m$. The parameter $m$ determines the duration of
942: the short-term regime for which the entropy increases
943: at a rate of 2 bits per iterations.
944:
945: Higher rates of entropy increase become possible for hierarchical
946: coarse-grainings, i.e.\ coarse-grained histories with coarse-graining on
947: several different scales of the phase space. As before we find approximate
948: decoherence for such sets of histories and a probability distribution which is
949: peaked at histories displaying regularities according to the classical shift
950: property. The following results are valid for large $l$
951: (classical limit) and values for $s_j$ $(j=1,2,\dots,\lambda)$ that are
952: significantly greater than $1$.
953: \begin{itemize}
954: \item Entropy after $k$ iteration steps in the case
955: $k<\mbox{min}\{m_1,m_2, \dots, m_{\lambda-1}\}$:
956: \begin{equation}
957: H[\{
958: h_{\vec{\ys}^1,\,\vec{\ys}^2,\dots,\vec{\ys}^{\lambda}}\}]
959: =
960: \lambda\cdot k\;+\;{\cal
961: O}(\frac{(l+r-k)\log_2(l+r-k)}{2^{l-2(k^2+(1+m_1+m_2+\dots+m_{\lambda-1})k)}})\;
962: \end{equation}
963: \end{itemize}
964: \begin{itemize}
965: \item Entropy after $k$ iteration steps in the case
966: $k>\mbox{max}\{m_1,m_2, \dots, m_{\lambda-1}\}$:
967: \begin{equation}
968: H[\{
969: h_{\vec{\ys}^1,\,\vec{\ys}^2,\dots,\vec{\ys}^{\lambda}}\}] =
970: k+\sum_{i=1}^{\lambda-1}m_i\;+\;{\cal
971: O}(\frac{(l+r-k)\log_2(l+r-k)}{2^{l-2(k^2+(1+m_1+m_2+\dots+m_{\lambda-1})k)}})\;.
972: \end{equation}
973: \end{itemize}
974: We see that in the long-term regime,
975: $k>\mbox{max}\{m_1,m_2, \dots, m_{\lambda-1}\}$,
976: the rate of entropy increase is again 1 bit per iteration, independently of
977: the character of the coarse-graining. In the short-term regime,
978: $k<\mbox{min}\{m_1,m_2, \dots, m_{\lambda-1}\}$, however,
979: the rate of entropy increase is $\lambda$ bits per iteration. The short-term
980: regime is thus characterised by $\lambda$, the number of coarse-graining
981: scales. The parameters $m_1,\ldots,m_{\lambda-1}$ determine the duration
982: of the short-term regime. Classical predictability decreases with increasing number
983: of coarse-graining scales.
984:
985: Finally, we note how the above results for the entropy
986: production in the various coarse-grained descriptions can be understood using
987: the shift property of the coarse-grained evolution of the quantum baker's map,
988: which is explained and illustrated in detail in the next section. For this we
989: make use of our diagram notation~(\ref{multi-scales_coarse-graining_diagram}).
990: The shift property implies that the only histories with significant
991: probabilities are those that satisfy the shift condition, i.e., the projectors
992: of the histories have to be related to the initial state via a shift. For
993: instance, if $ \rho_0 \propto {P}_{\xs^1,\,\xs^2,\dots,\xs^{\lambda}}^{(l,m_1,
994: m_2,\dots,m_{\lambda-1},r)}\;$, then only such histories can arise with
995: significant probabilities whose first event, represented by the projector
996: ${P}_{\ys^{1,1},\ys^{1,2},\dots, \ys^{1,\lambda}}^{(l,m_1,
997: m_2,\dots,m_{\lambda-1},r)}$, satisfies the shift constraint. Unless
998: $\y^{1,1}_{1:(s_1-1)}=\x^1_{2:s_1}$ and $\y^{1,2}_{1:(s_2-1)}=\x^2_{2:s_2}$
999: and $\dots$ and $\y^{1,\lambda}_{1:(s_{\lambda}-1)}=\x^\lambda_{2:s_\lambda}$
1000: is satisfied by the first event the whole history will have a vanishing
1001: probability. On the other hand the last bits $y^{1,1}_{s_1}$, $y^{1,2}_{s_2}$,
1002: $\dots$, $y^{1,\lambda}_{s_\lambda}$ of the strings $\y^{1,1}$, $\y^{1,2}$,
1003: $\dots$, $\y^{1,\lambda}$, which denote the first event of the history, remain
1004: undetermined, because the unspecified bits of the empty boxes
1005: in~(\ref{multi-scales_coarse-graining_diagram}), which are coarse-grained
1006: over, are shifted onto them. The bits $y^{1,1}_{s_1}$, $y^{1,2}_{s_2}$,
1007: $\dots$, $y^{1,\lambda}_{s_\lambda}$ may therefore be chosen arbitrarily,
1008: corresponding to a branching into $2^\lambda$ possible histories with
1009: non-vanishing probabilities. This branching into $2^\lambda$ alternatives
1010: repeats with each iteration step of the evolution, as long as
1011: $k<\mbox{min}\{m_1,m_2, \dots, m_{\lambda-1}\}$, leading to an entropy
1012: production of $\lambda$ bits per iteration step. As soon as the number of
1013: iterations $k$ starts to exceed, step by step, the values of $m_1,m_2, \dots,
1014: m_{\lambda-1}$, the rate of entropy production goes down, step by
1015: step, from the value $\lambda$ to the value $1$ in the long-term regime.
1016: Consider, for instance, the case in which $k>m_{\lambda-1}$. Only in the first
1017: $m_{\lambda-1}$ iteration steps coarse-grained bits (the empty boxes of our
1018: diagram notation) are shifted onto the last bits of the strings
1019: $\y^{j,\lambda-1}$, thereby making them arbitrarily
1020: chose-able for the history, causing a branching into two
1021: alternatives, and increasing the entropy by 1 bit. In the subsequent
1022: $k-m_{\lambda-1}$ iterations the string $\x^\lambda$ of the
1023: initial condition enters the scale of the $\y^{j,\lambda-1}$-strings, with the
1024: consequence that the last bits of the strings
1025: $\y^{m_{\lambda-1}+1,\lambda-1},\dots,\y^{k,\lambda-1}$ become determined by
1026: the initial condition, meaning no branching and therefore no entropy
1027: increase.
1028:
1029:
1030: \subsection{Derivation and illustration of the results}
1031: \label{sec:Derivations}
1032:
1033: \subsubsection{1-scale and 2-scale coarse-grainings}
1034: \label{sec:1-scale and 2-scale coarse-grainings}
1035:
1036: The decoherence functional for the locally
1037: coarse-grained histories (\ref{dfunc_local}) was
1038: calculated in an earlier work of two of the
1039: authors~\cite{Soklakov2002}. We briefly review
1040: the corresponding result, which is:
1041:
1042: \begin{equation} \label{dfunc_local_result}
1043: {\mathcal D}_{B,\,\rho^{\scriptscriptstyle (l,r)}_{\xs}}[h_{\vec{\ys}},
1044: h_{\vec{\zs}} ]
1045: =2^{-k}
1046: \underbrace{
1047: \left(\prod_{j=1}^{k}\delta_{\ys^j}^{\zs^j}\right)
1048: }_{{\rm diagonal}}
1049: \cdot
1050: \underbrace{\left(
1051: \delta{}_{\ys^1_{1:\gamma-1}}^{\xs_{2:\gamma}}
1052: \prod_{j=1}^{k-1}
1053: \delta{}_{\ys^{j+1}_{1:\gammas-1}}^{\;\ys^{j}_{2:\gammas}}
1054: \right)}_{{\rm step-by-step\ shift}}
1055: \cdot
1056: \underbrace{ \Bigg{(}
1057: \delta{}_{\ys^k_{1:\gamma-k}}^{\xs_{k+1:\gamma}}
1058: \Bigg{)} }_{k{\rm th\ shift }}
1059: \;\; +\,O(\frac{l+r-k}{2^{l-2(k^2+k)}})\;,
1060: \end{equation}
1061: where $\gamma\equiv |\x|=|\y^j|=|\z^j|=N-(l+r)$.
1062: The expression in the first parentheses is zero for all
1063: off-diagonal elements of the decoherence functional.
1064: In the limit of very large $l$ all off-diagonal elements
1065: of the decoherence functional vanish, the decoherence condition
1066: being therefore established. The diagonal elements of the decoherence
1067: functional can therefore be interpreted as probabilities of the corresponding
1068: histories (see Ref.~\cite{DowkerHalliwell1992} for a discussion of approximate
1069: decoherence). Asymptotically, only $2^{k}$ diagonal elements
1070: survive. Moreover, the error terms are exponentially small.
1071: We therefore get $2^{k}$ histories with asymptotically
1072: equal probabilities. The number of such histories doubles after
1073: each iteration step resulting in a loss of information
1074: at the rate of 1 bit per step. This information loss
1075: is quantified by the entropy increase of the set of histories.
1076: Since in the limit of large $l$ the set of histories
1077: $\{h_{\vec{\ys}}\}$ is decoherent, the individual alternative
1078: histories may be assigned probabilities, which are then given by
1079: $p[h_{\vec{\ys}}] = {\mathcal D}_{B,\,\rho^{\scriptscriptstyle
1080: (l,r)}_{\xs}}[h_{\vec{\ys}},h_{\vec{\ys}} ]$. Having found
1081: the probability distribution we may also define the entropy
1082: of the set of all possible alternative histories:
1083: \begin{eqnarray}
1084: H[\{h_{\vec{\ys}} \}]
1085: &\equiv& -\sum_{\vec{\ys}}p[h_{\vec{\ys}}]\log_2p[h_{\vec{\ys}}]\nonumber\\
1086: &\equiv& -\sum_{\vec{\ys}}{\mathcal D}_{B,\,\rho^{\scriptscriptstyle
1087: (l,r)}_{\xs}}[h_{\vec{\ys}},h_{\vec{\ys}} ]\log_2
1088: \left({\mathcal D}_{B,\,\rho^{\scriptscriptstyle
1089: (l,r)}_{\xs}}[h_{\vec{\ys}},h_{\vec{\ys}} ]\right)\;.
1090: \end{eqnarray}
1091: With (\ref{dfunc_local_result}) we obtain:
1092: \begin{equation} \label{EntropyLocal}
1093: H[\{h_{\vec{\ys}} \}] = k+
1094: O(\frac{(l+r-k)\log_2(l+r-k)}{2^{l-2(k^2+k)}})\,.
1095: \end{equation}
1096: In the limit of large $l$, for any fixed number of iterations, $k$, the
1097: entropy of the coarse-grained quantum baker's map approaches the value of $k$
1098: bits, i.e., 1 bit per iteration.
1099:
1100: What kind of histories arise with significant probabilities?
1101: This is determined by the expressions within the second and
1102: third parentheses of the result (\ref{dfunc_local_result}).
1103: Accordingly only histories that satisfy a
1104: {\em step-by-step shift condition\/}
1105: arise with significant probabilities.
1106: This can be illustrated using the diagram notation introduced above:
1107: \begin{eqnarray}
1108: & \underbrace{\Box\Box\dots\Box}_{l}\;
1109: \ \x_1
1110: \underline{\x_2\dots\x_{\gamma-2}\x_{\gamma-1} \x_\gamma } \;
1111: \underbrace{\Box\Box\dots \Box}_{r}\ , \cr
1112: & \ \ \ \ \ \ \boldarrow_{\phantom{|_{|_{|_|}}}} \cr
1113: & \underbrace{\Box\Box\dots\Box}_{l}\;
1114: \overline{ \ \y^1_1\underline{\y^1_2\dots\y^1_{\gamma-2} \y^1_{\gamma-1}}}
1115: \underline{\; y^1_\gamma}\;
1116: \underbrace{\Box\Box\dots \Box}_{r}\ , \cr
1117: & \ \ \ \ \ \ \boldarrow_{\phantom{|_{|_{|_|}}}} \cr
1118: & \underbrace{\Box\Box\dots\Box}_{l}\;
1119: \overline{ \ \y^2_1 \underline{\y^2_2\dots\,\y^2_{\gamma-2} y^2_{\gamma-1}}}
1120: \underline{\; y^2_\gamma}\;
1121: \underbrace{\Box\Box\dots \Box}_{r}\ , \cr
1122: & \ \ \ \ \boldarrow \cr
1123: & \ \ \ \ \dots \cr
1124: & \ \ \ \ \ \ \boldarrow_{\phantom{|_{|_{|_|}}}} \cr
1125: & \underbrace{\Box\Box\dots\Box}_{l}\;
1126: \overline{ \y^k_1 \dots\y^k_{\gamma-k} y^k_{\gamma-k+1}
1127: \!\dots}\, y^k_\gamma\,
1128: \underbrace{\Box\Box\dots \Box}_{r}\ .
1129: \end{eqnarray}
1130: The first line of this diagram represents the initial condition
1131: $\rho^{{\scriptscriptstyle{(l,r)}}}_{\xs}$. The subsequent lines
1132: correspond to the projectors $P^{(l,r)}_{\ys^1},\dots,P^{(l,r)}_{\ys^k}$
1133: constituting the history $h_{\vec{\ys}} $.
1134: The step-by-step shift condition is depicted by arrows and lines.
1135: Underlined substrings are shifted onto those overlined substrings
1136: which are indicated by arrows.
1137: In order to fulfil the step-by-step shift condition
1138: all underlined and overlined substrings that are connected by
1139: an arrow have to be equal. In this way it becomes clear which
1140: bits of the symbolic specification of a history are completely
1141: determined by the initial condition. These bits are in bold face.
1142: The other bits may be chosen arbitrarily. For instance, in the first iteration
1143: step the initial condition substring
1144: $\x_{2:\gamma}\equiv x_2\dots x_{\gamma}$ is
1145: shifted onto the substring $\y^{1}_{1:\gamma-1}
1146: \equiv y^1_1\dots y^1_{\gamma-1}$.
1147: The first $\gamma-1$ bits of the string $\y^1$ of the
1148: first event in the history $h_{\vec{\ys}} $ are therefore
1149: determined by the initial condition.
1150: Unless $\y^{1}_{1:\gamma-1}=\x_{2:\gamma}$ is satisfied by
1151: the first event, the whole history will have a vanishing
1152: probability. On the other hand the last bit $y^{1}_{\gamma}$
1153: of the string $\y^{1}$, which denotes the first event of the history,
1154: remains undetermined, because the unspecified bit of the empty box is
1155: shifted onto it, which is coarse-grained (i.e.\ summed) over.
1156: The bit $y^{1}_{\gamma}$ may therefore be chosen arbitrarily, corresponding
1157: to a branching into two possible histories with non-vanishing
1158: probabilities and therefore an entropy increase of 1 bit. This
1159: procedure repeats with each iteration step of the evolution.
1160: For the entire history, therefore, there are only $k$ independent
1161: bits which can be chosen arbitrarily, given the step-by-step shift
1162: constraint.
1163:
1164: The calculation of the decoherence functional (\ref{checkerboard-dfunc})
1165: for the non-locally coarse-grained histories can be traced back to using ´
1166: the above result for the local ones. To do so, we may express
1167: all the nonlocal projection operators appearing in the
1168: decoherence functional as sums over suitable local ones:
1169: \begin{eqnarray}
1170: \rho_{\xs^1,\,\xs^2}^{(l,m_l, m_r, r)}&=& 2^{-(l+m_l+m_r+r)}
1171: \sum_{\,\bxis\in \{0,1\}^{m_l+m_r}} {P}_{\xs^1\bxis\xs^2}^{(l,r)}\;,\\
1172: {P}_{\ys^{j,1},\,\ys^{j,2}}^{(l,m_l, m_r, r)}&=&
1173: \sum_{\,\boldetas^j\in \{0,1\}^{m_l+m_r}}
1174: {P}_{\ys^{j,1}\boldetas^j\ys^{j,2}}^{(l,r)}\;\;,\;\; j=1,2,\dots,k\;,\\
1175: {P}_{\zs^{j,1},\,\zs^{j,2}}^{(l,m_l, m_r, r)}&=&
1176: \sum_{\,\bzetas^j\in \{0,1\}^{m_l+m_r}}
1177: {P}_{\zs^{j,1}\bzetas^j\zs^{j,2}}^{(l,r)}\;\;,\;\;\; j=1,2,\dots,k\;.
1178: \end{eqnarray}
1179: By inserting these expressions into the decoherence functional
1180: (\ref{checkerboard-dfunc}) we arrive at:
1181: \begin{eqnarray}
1182: &&\hspace*{-1.3cm}
1183: {\mathcal D}_{B,\,\rho_0}[h_{\vec{\ys}^1,\,\vec{\ys}^2},
1184: h_{\vec{\zs}^1,\,\vec{\zs}^2} ]=\nonumber\\&=&
1185: \sum_{\,\boldetas^1\in \{0,1\}^{m_l+m_r}}\dots
1186: \sum_{\,\boldetas^k\in \{0,1\}^{m_l+m_r}}
1187: \sum_{\,\bxis\in \{0,1\}^{m_l+m_r}}
1188: \sum_{\,\bzetas^1\in \{0,1\}^{m_l+m_r}}\dots
1189: \sum_{\,\bzetas^k\in \{0,1\}^{m_l+m_r}}
1190: \nonumber\\&&
1191: 2^{-(m_l+m_r)}\Tr\Big[ \phantom{\Big]}
1192: {P}_{\ys^{k,1}\boldetas^k\ys^{k,2}}^{(l,r)}B
1193: {P}_{\ys^{k-1,1}\boldetas^{k-1}\ys^{k-1,2}}^{(l,r)}
1194: B\cdots B
1195: {P}_{\ys^{1,1}\boldetas^1\ys^{1,2}}^{(l,r)}\times \nonumber \\
1196: &&\times \phantom{\Big[} B
1197: \big(\frac{1}{2^{l+r}}{P}_{\xs^1\xibolds\xs^2}^{(l,r)} \big)
1198: B^\dag {P}_{\zs^{1,1}\bzetas^1\zs^{1,2}}^{(l,r)}
1199: B^\dag\cdots
1200: {P}_{\zs^{k-1,1}\bzetas^{k-1}\zs^{k-1,2}}^{(l,r)}B^\dag
1201: {P}_{\zs^{k,1}\bzetas^{k}\zs^{k,2}}^{(l,r)}\Big]\nonumber\\&&
1202: \end{eqnarray}
1203: Each term of the sum over all possible strings $\bxi$, $\{\boldeta^j\}$
1204: and $\{\bzeta^j\}$ is, apart from the factor $2^{-(m_l+m_r)}$, a
1205: decoherence functional with respect to histories composed of
1206: local projectors. Each such term, therefore, results in an expression
1207: of the form (\ref{dfunc_local_result}), and we obtain:
1208: \begin{eqnarray}
1209: &&\hspace*{-1.3cm}
1210: {\mathcal D}_{B,\,\rho_0}[h_{\vec{\ys}^1,\,\vec{\ys}^2},
1211: h_{\vec{\zs}^1,\,\vec{\zs}^2} ]=\nonumber\\&=&
1212: \sum_{\,\boldetas^1\in \{0,1\}^{m_l+m_r}}\dots
1213: \sum_{\,\boldetas^k\in \{0,1\}^{m_l+m_r}}
1214: \sum_{\,\bxis\in \{0,1\}^{m_l+m_r}}
1215: \sum_{\,\bzetas^1\in \{0,1\}^{m_l+m_r}}\dots
1216: \sum_{\,\bzetas^k\in \{0,1\}^{m_l+m_r}}2^{-(m_l+m_r)}\times\nonumber\\&&
1217: \times\Bigg\{ \phantom{\Bigg\} } 2^{-k}
1218: \underbrace{
1219: \left(\prod_{i=1}^{k}\delta_{\ys^{i,1}\boldetas^i\ys^{i,2}}
1220: ^{\zs^{i,1}\bzetas^i\zs^{i,2}}\right)
1221: }_{{\rm diagonal}}\cdot
1222: \underbrace{\left(
1223: \delta{}_{\ys^{1,1}\boldetas^1\ys^{1,2}_{1:s_2-1}}
1224: ^{\xs^1_{2:s_1}\bxis\,\xs^2}
1225: \prod_{j=1}^{k-1}
1226: \delta{}_{\ys^{j+1,1}\boldetas^{j+1}\ys^{j+1,2}_{1:(s_2-1)}}
1227: ^{\ys^{j,1}_{2:s_1}\boldetas^{j}\ys^{j,2}}\right)}_{{\rm step-by-step\
1228: shift}}\times
1229: \nonumber\\&&\times \phantom{\Bigg\{ }
1230: \underbrace{ \Bigg{(}
1231: \delta{}_{(\ys^{k,1}\boldetas^k\ys^{k,2})_{1:\gamma-k}}
1232: ^{(\xs^1\bxis\,\xs^2)_{k+1:\gamma}}
1233: \Bigg{)} }_{k{\rm th\ shift }}
1234: \;\; +\,{\cal O}(\frac{l+r-k}{2^{l-2(k^2+k)}})\;
1235: \Bigg\}\;\nonumber\\
1236: &=&
1237: \sum_{\,\boldetas^1\in \{0,1\}^{m_l+m_r}}\dots
1238: \sum_{\,\boldetas^k\in \{0,1\}^{m_l+m_r}}
1239: \sum_{\,\bxis\in \{0,1\}^{m_l+m_r}}\Bigg\{ \phantom{\Bigg\} }
1240: 2^{-(m_l+m_r)}\cdot 2^{-k}\cdot\underbrace{
1241: \left(\prod_{i=1}^{k}\delta_{\ys^{i,1}\boldetas^i\ys^{i,2}}
1242: ^{\zs^{i,1}\boldetas^i\zs^{i,2}}\right)
1243: }_{{\rm diagonal}}\times\nonumber\\ &&\times
1244: \underbrace{\left(
1245: \delta{}_{\ys^{1,1}}^{\xs^1_{2:s_1}\xi_1}
1246: \delta{}_{\boldetas^1}^{\bxis_{2:(m_l+m_r)}\xs^2_1}
1247: \delta{}_{\ys^{1,2}_{1:s_2-1}}^{\xs^2_{2:s_2}}\cdot
1248: \prod_{j=1}^{k-1}
1249: \delta{}_{\ys^{j+1,1}}^{\ys^{j,1}_{2:s_1}\eta^j_1}
1250: \delta{}_{\boldetas^{j+1}}^{\boldetas^{j}_{2:(m_l+m_r)}\ys^{j,2}_1}
1251: \delta{}_{\ys^{j+1,2}_{1:s_2-1}}^{\ys^{j,2}_{2:s_2}}
1252: \right)}_{{\rm step-by-step\ shift}}\times \nonumber\\&&\times
1253: \underbrace{ \Bigg{(}
1254: \delta{}_{(\ys^{k,1}\boldetas^k\ys^{k,2})_{1:\gamma-k}}
1255: ^{(\xs^1\bxis\,\xs^2)_{k+1:\gamma}}
1256: \Bigg{)} }_{k{\rm th\ shift }} \phantom{\Bigg\{ } \Bigg\}
1257: \;\; +\,\,\Big( 2^{m_l+m_r}\Big)^{2k}\cdot\,
1258: {\cal O}(\frac{l+r-k}{2^{l-2(k^2+k)}})\;
1259: \nonumber\\
1260: \end{eqnarray}
1261: Here $\gamma$ denotes the length of the strings
1262: $\y^{j,1}\boldeta^j\y^{j,2}$ and
1263: $\x^1\bxi\,\x^2$, respectively,
1264: i.e.\ $\gamma=|\x^1\bxi\,\x^2|=|\y^{j,1}\boldeta^j\y^{j,2}|=
1265: s_1+(m_l+m_r)+s_2$.
1266:
1267: First of all the sum over all possible $\bzeta^j\in
1268: \{0,1\}^{m_l+m_r}$, $j=1,\dots, k$, collapses due to the term
1269: $\prod_{i=1}^{k}\delta_{\ys^{i,1}\boldetas^i\ys^{i,2}}^
1270: {\zs^{i,1}\bzetas^i\zs^{i,2}}\,$, apart from contributing a
1271: factor $2^{k(m_l+m_r)}$ to the error term. Secondly we note
1272: that the step-by-step shift condition causes the whole sum
1273: $\sum_{\boldetas^1}\sum_{\boldetas^2}\dots\sum_{\boldetas^k}$
1274: to collapse, apart from contributing a further factor
1275: $2^{k(m_l+m_r)}$ to the bound on the error term,
1276: which is furthermore enlarged by a factor $2^{(m_l+m_r)}$
1277: stemming from the sum $\sum_{\bxis}$. Let us comprehend the
1278: collapse of the sums $\sum_{\boldetas^j}$. For a given fixed string
1279: $\bxi$ out of the sum $\sum_{\bxis}$ all $\boldeta^1, \boldeta^2,
1280: \dots, \boldeta^k$ are through the $\delta$'s determined by the
1281: string $\bxi$ and the given fixed string $\x^2$ of the initial
1282: condition. The first shift leads to a determination of $\boldeta^1$:
1283: according to $\delta{}_{\boldetas^1}^{\bxis_{2:(m_l+m_r)}\xs^2_1}$
1284: the sum over all possible $\boldeta^1\in\{0,1\}^{m_l+m_r}$ collapses
1285: and only the string $\boldeta^1=\hspace{-3.3mm}^!\;\;
1286: \bxi_{2:(m_l+m_r)}\x^2_1$ survives. The second shift determines
1287: $\boldeta^2$, since according to $\delta{}_{\boldetas^{2}}^
1288: {\boldetas^{1}_{2:(m_l+m_r)}\ys^{1,2}_1}$ the sum over all possible
1289: $\boldeta^2\in\{0,1\}^{m_l+m_r}$ collapses and only the string
1290: $\boldeta^2= \;\boldeta^1_{2:(m_l+m_r)}\y^{1,2}_1
1291: \equiv\bxi_{3:(m_l+m_r)}\x^2_1\x^2_2$ does lead to a non-vanishing
1292: contribution to the decoherence functional. It is easy to see that
1293: due to the step-by-step shift condition all the sums
1294: $\sum_{\boldetas^j}$, $j=1,\dots, k$, collapse and only the strings
1295: \begin{eqnarray}\label{determinationetas}
1296: \boldeta^j=\bxi_{(j+1):(m_l+m_r)}\x^2_{1:j}
1297: \end{eqnarray}
1298: out of these sums survive leading together to a non-vanishing
1299: contribution to the decoherence functional. In fact the step-by-step
1300: shift condition can also be expressed in the following way:
1301: \begin{equation}
1302: \prod_{j=1}^{k} \delta{}_{(\ys^{j,1}\boldetas^j\ys^{j,2})_{1:\gamma-j}}
1303: ^{(\xs^1\bxis\,\xs^2)_{j+1:\gamma}}\;\:,
1304: \end{equation}
1305: meaning that only such strings $\boldeta^j$ out of the
1306: corresponding sums $\sum_{\boldetas^j}$, $j=1,\dots, k$,
1307: lead to a non-vanishing contribution to the decoherence functional
1308: which are determined by $\bxi$ and $\x^2$ according to (\ref{determinationetas}).
1309: Next we note that as a consequence of the step-by-step shift condition
1310: also the sum over all possible $\bxi\in\{0,1\}^{m_l+m_r}$ collapses.
1311: It collapses only {\em partially} in case $k< m_l+m_r$ and it
1312: collapses {\em completely} in case $k \ge m_l+m_r$. Let us first consider
1313: the case $k< m_l+m_r$. After the first shift the first bit of $\bxi$
1314: is determined by the last bit of the string $\y^{1,1}$ of the given
1315: history, i.e.\ $\xi_1=\hspace{-3.4mm}^!\;\;y^{1,1}_{s_1}$,
1316: according to the term $\delta{}_{\ys^{1,1}}^{\xs^1_{2:s_1}\xi_1}$.
1317: The second shift leads to
1318: $\delta{}_{\ys^{2,1}}^{\ys^{1,1}_{2:s_1}\eta^1_1}$,
1319: so that $\eta_1^1= \;y^{2,1}_{s_1}$.
1320: But we have $\eta_1^1= \;\xi_2$ due to the
1321: first shift, so we arrive at a determination of $\xi_2$,
1322: namely $\xi_2= \;y^{2,1}_{s_1}$. In this
1323: way the sum over all possible $\bxi\in\{0,1\}^{m_l+m_r}$ collapses
1324: to a sum over all possible
1325: $\bxi_{(k+1):(m_l+m_r)}\in\{0,1\}^{m_l+m_r-k}$,
1326: \begin{equation}
1327: \sum_{\,\bxis\in \{0,1\}^{m_l+m_r}}\longrightarrow
1328: \sum_{\,\bxis_{(k+1):(m_l+m_r)}\in\{0,1\}^{m_l+m_r-k}}\;,
1329: \end{equation}
1330: since the first $k$ bits $\xi_1, \dots, \xi_k$ out of the sum $\sum_{\bxis}$
1331: have to fulfil the step-by-step shift condition and are therefore
1332: determined by $\xi_j= \;y^{j,1}_{s_1}$.
1333: That the first $k$ bits of the string $\bxi$ out of the sum
1334: $\sum_{\bxis}$ are determined by the given history
1335: $h_{\vec{\ys}^1,\,\vec{\ys}^2}$ and therefore the sum over
1336: the first $k$ bits of $\bxi= \xi_1\xi_2\dots\xi_{m_l+m_r}$
1337: collapses can also be seen by looking at the $k$-th shift factor
1338: which in fact appears as a redundant factor in the result: according
1339: to $ \delta{}_{(\ys^{k,1}\boldetas^k\ys^{k,2})_{1:\gamma-k}}
1340: ^{(\xs^1\bxis\,\xs^2)_{k+1:\gamma}}$ only such strings $\bxi$
1341: out of the sum $\sum_{\bxis}$ lead to a non-vanishing contribution to
1342: the decoherence functional for which $\bxi_{1:k}= \;
1343: \y^{k,1}_{(s_1-k+1):s_1}$ holds. The remaining $m_l+m_r-k$ bits of
1344: $\bxi= \xi_1\xi_2\dots\xi_{m_l+m_r}$ remain undetermined and are
1345: still summed over. There are $2^{m_l+m_r-k}$ possible different
1346: substrings $\bxi_{k+1:m_l+m_r}\in\{0,1\}^{m_l+m_r-k}$ in this
1347: remaining sum leading to a non-vanishing contribution to the
1348: decoherence functional. Since the contributions of all these strings
1349: are equal, as can be seen by looking at the result, we may replace
1350: the remaining sum over all possible $\bxi_{k+1:m_l+m_r}$ by the factor
1351: $2^{m_l+m_r-k}$. Furthermore all the $\delta$-terms containing bits of the
1352: unspecified strings $\bxi$ and $\boldeta^j$, $j=1,\dots,k$, which are
1353: summed over, may now be replaced by $1$ after having been exploited
1354: for the determination of that strings $\bxi$ and $\boldeta^j$ out
1355: of the sums $\sum_{\bxis}$ and $\sum_{\boldetas^1}\sum_{\boldetas^2}
1356: \dots\sum_{\boldetas^k}$ which lead to a non-vanishing contribution
1357: to the value of the decoherence functional. In case $k<m_l+m_r$ we
1358: therefore arrive at the following result:
1359:
1360: \begin{itemize}
1361: \item Decoherence functional in case $k<m_l+m_r$:
1362: \begin{eqnarray}\label{result_k<(m_l+m_r)}
1363: {\mathcal D}_{B,\,\rho_0}[h_{\vec{\ys}^1,\,\vec{\ys}^2},
1364: h_{\vec{\zs}^1,\,\vec{\zs}^2} ]&=&
1365: \underbrace{2^{m_l+m_r-k}\cdot 2^{-(m_l+m_r)}\cdot 2^{-k}}_{=\,2^{-2k}}
1366: \cdot\underbrace{\left(\prod_{i=1}^{k}\delta_{\ys^{i,1}}^{\zs^{i,1}}
1367: \delta_{\ys^{i,2}}^{\zs^{i,2}}\right)}_{{\rm diagonal}}\times\nonumber\\
1368: &&\times
1369: \underbrace{\left(
1370: \delta{}_{\ys^{1,1}_{1:s_1-1}}^{\xs^1_{2:s_1}}
1371: \delta{}_{\ys^{1,2}_{1:s_2-1}}^{\xs^2_{2:s_2}}\cdot
1372: \prod_{j=1}^{k-1}
1373: \delta{}_{\ys^{j+1,1}_{1:s_1-1}}^{\ys^{j,1}_{2:s_1}}
1374: \delta{}_{\ys^{j+1,2}_{1:s_2-1}}^{\ys^{j,2}_{2:s_2}}
1375: \right)}_{{\rm step-by-step\ shift}}\times \\&&\times
1376: \underbrace{ \Bigg{(}
1377: \delta{}_{\ys^{k,1}_{1:s_1-k}}^{\xs^1_{k+1:s_1}}
1378: \delta{}_{\ys^{k,2}_{1:s_2-k}}^{\xs^2_{k+1:s_2}}
1379: \Bigg{)} }_{k{\rm th\ shift }}
1380: \;\; +\,\, {\cal
1381: O}(\frac{l+r-k}{2^{l-2(k^2+(1+m_l+m_r)k)}})\;.
1382: \nonumber
1383: \end{eqnarray}
1384: \end{itemize}
1385:
1386: \noindent
1387: Let us now consider the case $k\ge m_l+m_r$. As already mentioned in
1388: this case the whole sum $\sum_{\bxis}$ collapses to a single string
1389: $\bxi\in\{0,1\}^{m_l+m_r}$ satisfying the step-by-step shift
1390: condition. This can be seen, again, by looking at the $k$-th shift
1391: condition given by the factor $\delta{}_{(\ys^{k,1}\boldetas^k
1392: \ys^{k,2})_{1:\gamma-k}}^{(\xs^1\bxis\,\xs^2)_{k+1:\gamma}}$;
1393: according to it each string $\bxi$ out of the sum $\sum_{\bxis}$
1394: is shifted onto $(m_l+m_r)$ bits of the string $\y^{k,1}$, but since
1395: $\y^{k,1}$ is a {\em fixed} string specifying the last event of the
1396: {\em given} history, only the string
1397: $\bxi= \;\y^{k,1}_{(s_1-k+1):(s_1-k+m_l+m_r)}$
1398: out of the sum $\sum_{\bxis}$ survives. Of course we presupposed,
1399: or had to require, that $s_1\ge k\ge m_l+m_r$.
1400: In case $k\ge m_l+m_r$ we therefore get:
1401:
1402:
1403: \begin{itemize}
1404: \item Decoherence functional in case $k\ge m_l+m_r$:
1405: \begin{eqnarray} \label{result_k>(m_l+m_r)}
1406: {\mathcal D}_{B,\,\rho_0}[h_{\vec{\ys}^1,\,\vec{\ys}^2},
1407: h_{\vec{\zs}^1,\,\vec{\zs}^2} ]&=&
1408: 2^{-(m_l+m_r)} \cdot 2^{-k}
1409: \cdot\underbrace{\left(\prod_{i=1}^{k}\delta_{\ys^{i,1}}^{\zs^{i,1}}
1410: \delta_{\ys^{i,2}}^{\zs^{i,2}}\right)}_{{\rm
1411: diagonal}}\times\nonumber\\&&
1412: \times
1413: \underbrace{\left(
1414: \delta{}_{\ys^{1,1}_{1:s_1-1}}^{\xs^1_{2:s_1}}
1415: \delta{}_{\ys^{1,2}_{1:s_2-1}}^{\xs^2_{2:s_2}}\cdot
1416: \prod_{j=1}^{k-1}
1417: \delta{}_{\ys^{j+1,1}_{1:s_1-1}}^{\ys^{j,1}_{2:s_1}}
1418: \delta{}_{\ys^{j+1,2}_{1:s_2-1}}^{\ys^{j,2}_{2:s_2}}
1419: \right)}_{{\rm step-by-step\ shift}}\times \\&&\times
1420: \underbrace{ \Bigg{(}
1421: \delta{}_{\ys^{k,1}_{1:s_1-k}}^{\xs^1_{k+1:s_1}}
1422: \delta{}_{\ys^{k,1}_{s_1-k+(m_l+m_r)+1\,:\,s_1}}^{\xs^2_{1:\,k-(m_l+m_r)}}
1423: \delta{}_{\ys^{k,2}_{1:s_2-k}}^{\xs^2_{k+1:s_2}}
1424: \Bigg{)} }_{k{\rm th\ shift }}\nonumber\\
1425: &&\;\; +\,\, {\cal
1426: O}(\frac{l+r-k}{2^{l-2(k^2+(1+m_l+m_r)k)}})\;.
1427: \nonumber
1428: \end{eqnarray}
1429: \end{itemize}
1430:
1431:
1432: Let us now discuss the results (\ref{result_k<(m_l+m_r)})
1433: and (\ref{result_k>(m_l+m_r)}) for the decoherence functional
1434: (\ref{checkerboard-dfunc}). First of all we get approximate
1435: decoherence: for very large $l$ the decoherence functional is
1436: approximately diagonal. In the asymptotic limit
1437: $l \rightarrow \infty $ our set of histories
1438: $\{ h_{\vec{\ys}^1,\,\vec{\ys}^2}\}$ becomes decoherent.
1439: The diagonal elements of the functional,
1440: $\mathcal{D}_{B,\,\rho_0}[h_{\vec{\ys}^1,\,\vec{\ys}^2},
1441: h_{\vec{\ys}^1,\,\vec{\ys}^2} ]$, may therefore be interpreted
1442: as probabilities of the corresponding histories, i.e.\
1443: $p(h_{\vec{\ys}^1,\,\vec{\ys}^2},)=
1444: \mathcal{D}_{B,\,\rho_0}[h_{\vec{\ys}^1,\,\vec{\ys}^2},
1445: h_{\vec{\ys}^1,\,\vec{\ys}^2} ]$.
1446: Again there is no single dominant history.
1447: Several different histories arise with significant probabilities$\,$.
1448: In case $k<m_l+m_r$ we get $2^{2k}$ different histories with
1449: asymptotically equal probabilities (given by $2^{-2k}$). The number
1450: of histories with asymptotically nonzero probabilities becomes
1451: four times larger after each iteration step of the quantum
1452: baker's map resulting in a {\em loss of information of $2$ bits per step}.
1453: The entropy increase is therefore {\em $2$ bits per iteration step},
1454: which can also be seen by calculating the entropy of the
1455: approximately decoherent set of histories $\{ h_{\vec{\ys}^1,\,\vec{\ys}^2}\}$:
1456:
1457: \begin{itemize}
1458: \item Entropy after $k$ iteration steps in case $k\le m_l+m_r$:
1459: \begin{eqnarray}
1460: H[\{ h_{\vec{\ys}^1,\,\vec{\ys}^2}\}]
1461: & = & -\sum_{\vec{\ys}^1,\,\vec{\ys}^2 }p[ h_{\vec{\ys}^1,\,\vec{\ys}^2}]
1462: \log_2p[h_{\vec{\ys}^1,\,\vec{\ys}^2}]\nonumber\\
1463: &=&
1464: 2k\;+\;{\cal
1465: O}(\frac{(l+r-k)\log_2(l+r-k)}{2^{l-2(k^2+(1+m_l+m_r)k)}})\;.
1466: \end{eqnarray}
1467: \end{itemize}
1468:
1469: \noindent
1470: Again, only such histories are allowed to arise with significant
1471: probabilities that satisfy the shift condition: the projectors
1472: of the histories have to be related to the initial state via a shift.
1473: Let us illustrate this issue once again by means of our diagram
1474: notation:
1475:
1476: \begin{eqnarray}
1477: & \underbrace{\Box\Box\dots\Box}_{l}\;
1478: \ \x_1^1
1479: \underline{\x^1_2\dots\x^1_{s_1-2}\x^1_{s_1-1}
1480: \x^1_{s_1}}\;\;\underbrace{\Box\Box\dots\Box}_{m_l+m_r} \;\ \x^2_1
1481: \underline{\x^2_2\dots\x^2_{s_2-2}\x^2_{s_2-1}\x^2_{s_2}} \;
1482: \underbrace{\Box\Box\dots \Box}_{r}\ \cr
1483: &\hspace{-0.5cm} \ \ \ \ \ \ \boldarrow_{\phantom{|_{|_{|_|}}}}
1484: \hspace{4.5cm}\ \ \ \ \ \ \boldarrow_{\phantom{|_{|_{|_|}}}}
1485: \cr
1486: & \underbrace{\Box\Box\dots\Box}_{l}\;
1487: \overline{ \
1488: \y^{1,1}_1\underline{\y^{1,1}_2\dots\y^{1,1}_{s_1-2}
1489: \y^{1,1}_{s_1-1}}} \underline{\; y^{1,1}_{s_1}}\;\;
1490: \underbrace{\Box\Box\dots\Box}_{m_l+m_r} \;\ \overline{ \
1491: \y^{1,2}_1\underline{\y^{1,2}_2\dots\y^{1,2}_{s_2-2}
1492: \y^{1,2}_{s_2-1}}} \underline{\; y^{1,2}_{s_2}}\;\;
1493: \underbrace{\Box\Box\dots \Box}_{r}\ \cr
1494: &\hspace{-0.5cm} \ \ \ \ \ \ \boldarrow_{\phantom{|_{|_{|_|}}}}
1495: \hspace{5.5cm}\ \ \ \ \ \ \boldarrow_{\phantom{|_{|_{|_|}}}}\cr
1496: & \underbrace{\Box\Box\dots\Box}_{l}\;
1497: \overline{ \
1498: \y^{2,1}_1\underline{\y^{2,1}_2\dots\y^{2,1}_{s_1-2}
1499: y^{2,1}_{s_1-1}}} \underline{\; y^{2,1}_{s_1}}\;\;
1500: \underbrace{\Box\Box\dots\Box}_{m_l+m_r} \;\ \overline{ \
1501: \y^{2,2}_1\underline{\y^{2,2}_2\dots\y^{2,2}_{s_2-2}
1502: y^{2,2}_{s_2-1}}} \underline{\; y^{2,2}_{s_2}}\;\;
1503: \underbrace{\Box\Box\dots \Box}_{r}\ \cr
1504: &\hspace{-0.5cm} \ \ \ \ \ \ \boldarrow_{\phantom{|_{|_{|_|}}}}
1505: \hspace{5.5cm}\ \ \ \ \ \ \boldarrow_{\phantom{|_{|_{|_|}}}}\cr
1506: & \hspace{-0.5cm}\ \ \ \ \dots\hspace{6cm}\ \ \ \ \ \dots \cr
1507: &\hspace{-0.5cm} \ \ \ \ \ \ \boldarrow_{\phantom{|_{|_{|_|}}}}
1508: \hspace{5.5cm}\ \ \ \ \ \ \boldarrow_{\phantom{|_{|_{|_|}}}}\cr
1509: & \underbrace{\Box\Box\dots\Box}_{l}\;
1510: \overline{ \y^{k,1}_1 \dots\y^{k,1}_{s_1-k} y^{k,1}_{s_1-k+1}
1511: \!\dots}\, y^{k,1}_{s_1}\;\;
1512: \underbrace{\Box\Box\dots\Box}_{m_l+m_r} \;\
1513: \overline{ \y^{k,2}_1 \dots\y^{k,2}_{s_2-k} y^{k,2}_{s_2-k+1}
1514: \!\dots}\, y^{k,2}_{s_2}\;\;
1515: \underbrace{\Box\Box\dots \Box}_{r}\ \cr \nonumber
1516: \end{eqnarray}
1517: \begin{equation}
1518: \end{equation}
1519:
1520: This diagram illustrates symbolically the content of the result
1521: (\ref{result_k<(m_l+m_r)}). Again, the first line of this diagram represents
1522: the initial condition $\rho_0 = \rho_{\xs^1,\,\xs^2}^{{\scriptstyle{(l,m_l,
1523: m_r, r)}}}$. The subsequent lines correspond to the projectors
1524: ${P}_{\ys^{1,1},\,\ys^{1,2}}^{(l,m_l, m_r, r)},\dots,
1525: {P}_{\ys^{k,1},\,\ys^{k,2}}^{(l,m_l, m_r, r)}$ representing the subsequent
1526: propositions of the history $ h_{\vec{\ys}^1,\,\vec{\ys}^2}$. The
1527: coarse-grained islands in the middle of each line, with $m_l+m_r$ empty boxes
1528: each, subsequently represent the sums $\sum_{\bxis}$, $\sum_{\boldetas^1},
1529: \sum_{\boldetas^2},\dots\,\sum_{\boldetas^k}$ in our calculation. Again, the
1530: step-by-step shift condition is depicted by arrows and lines. Underlined
1531: substrings are shifted onto those overlined substrings which are indicated by
1532: arrows. In order to fulfil the step-by-step shift condition all underlined
1533: and overlined substrings that are connected by an arrow must be equal. In this
1534: way we immediately see which bits of the symbolic specification of a history
1535: are completely determined by the initial condition. In the diagram these bits
1536: are indicated by using bold face. The remaining bits, which are not in bold
1537: face, may be chosen arbitrarily. For instance, in the first iteration step the
1538: initial condition substrings $\;\x^1_{2:s_1}\equiv x^1_2\dots x^1_{s_1-2}
1539: x^1_{s_1-1} x^1_{s_1}\;$ and $\;\x^2_{2:s_2}\equiv x^2_2\dots x^2_{s_2-2}
1540: x^2_{s_2-1} x^2_{s_2}\;$ are shifted onto the substrings
1541: $\;\y^{1,1}_{1:(s_1-1)}\equiv y^{1,1}_1 y^{1,1}_2\dots
1542: y^{1,1}_{s_1-2}y^{1,1}_{s_1-1}\;$ and $\;\y^{1,2}_{1:(s_2-1)}\equiv y^{2,2}_1
1543: y^{2,2}_2\dots y^{2,2}_{s_2-2}y^{2,2}_{s_2-1}\;$, respectively. The first
1544: $(s_1-1)$ bits of the string $\y^{1,1}$ and the first $(s_2-1)$ bits of the
1545: string $\y^{1,2}$ of the first event in the history are therefore determined
1546: by the initial condition. Unless $\y^{1,1}_{1:(s_1-1)}=\x^1_{2:s_1}$ and
1547: $\y^{1,2}_{1:(s_2-1)}=\x^2_{2:s_2}$ is satisfied by the first event the whole
1548: history will have vanishing probability. On the other hand the last bits
1549: $y^{1,1}_{s_1}$ and $y^{1,2}_{s_2}$ of the strings $\y^{1,1}$ and $\y^{1,2}$,
1550: which denote the first event of the history, remain undetermined, because the
1551: unspecified bits of the empty boxes are shifted onto them, which are
1552: coarse-grained (i.e.\ summed) over. The bits $y^{1,1}_{s_1}$ and
1553: $y^{1,2}_{s_2}$ may therefore be chosen arbitrarily resulting in a branching into
1554: four possible histories with non-vanishing probabilities. This procedure
1555: repeats with each iteration step of the evolution. The second step leads to a
1556: determination of the first $(s_1-1)$ bits of the string $\y^{2,1}$ and the
1557: first $(s_2-1)$ bits of the string $\y^{2,2}$ symbolising the second event of
1558: the history, whereas, again, the last bits of these strings remain unspecified
1559: and may be chosen arbitrarily implicating a branching into further four
1560: alternatives with non-vanishing probabilities. And so on. It becomes clear
1561: from the above picture which histories arise with significant probabilities
1562: during the evolution and why the number of alternative equiprobable histories
1563: is quadruplicated after each iteration step. After $k$ iteration steps---we
1564: still assume $k<m_l+m_r$---there are therefore $2k$ independent bits which can
1565: be chosen arbitrarily, given the step-by-step shift constraint. This
1566: implicates $2^{2k}$ alternative, equiprobable histories that may arise with
1567: significant probability after $k$ iteration steps.
1568:
1569:
1570: Our result for $k>m_l+m_r$, Eq.\ (\ref{result_k>(m_l+m_r)}),
1571: may be interpreted in the following way.
1572: As long as the number of iterations $k$ is smaller than $m=m_l+m_r$
1573: the number of histories with asymptotically non-vanishing
1574: probabilities becomes four times larger after each iteration step of
1575: the quantum baker's map resulting in an entropy increase
1576: of $2$ bits per iteration
1577: step. As soon as the number of iterations becomes greater than
1578: $m=m_l+m_r$, the entropy increase becomes $1$ bit per iteration
1579: step. This is what is expressed by the result $2^{-(m_l+m_r)}\cdot
1580: 2^{-k}= 2^{-2(m_l+m_r)}\cdot 2^{-(k-(m_l+m_r))}$ for the probability
1581: of the histories which are allowed to occur. The first
1582: $m_l+m_r$ iteration steps lead to an entropy increase of $2$ bits
1583: per step involving $2^{2(m_l+m_r)}$ asymptotically equiprobable
1584: histories. The remaining $k-(m_l+m_r)$ iteration steps produce
1585: an entropy increase of $1$ bit per step only, with the number of
1586: histories with significant probabilities being doubled at each step,
1587: implicating a branching factor $2^{k-(m_l+m_r)}$. The entire number
1588: of histories arising with significant probabilities after
1589: $k$ iteration steps therefore becomes $2^{2(m_l+m_r)}\cdot 2^{k-(m_l+m_r)}
1590: =2^{(m_l+m_r)}\cdot 2^k$, the histories being asymptotically
1591: equiprobable. Again the issue becomes clearer when using our diagram
1592: picture. The size of the middle coarse-grained islands
1593: is now only $m_l+m_r<k$. So only in the first $m_l+m_r$ iteration
1594: steps coarse-grained bits are shifted onto the last bits
1595: of the strings $\y^{j,1}$, making them by this means unspecified, i.e.\
1596: arbitrarily chose-able for the history. In the subsequent,
1597: remaining $k-(m_l+m_r)$ iteration steps the string $\x^2$ of the
1598: initial condition enters the scale of the $\y^{j,1}$-strings,
1599: with the consequence that the last bits of the strings
1600: $\y^{m_l+m_r+1,1},\dots,\y^{k,1}$
1601: become determined by the initial condition.
1602: At the end, after the $k$-th iteration step, only $m_l+m_r$ bits
1603: of the string $\y^{k,1}$ may be chosen arbitrarily, the first
1604: $s_1-k$ bits and the last $k-(m_l+m_r)$ bits of it being determined
1605: by the initial condition. On the other hand only the first $s_2-k$
1606: bits of the string $\y^{k,2}$ become determined by the initial
1607: condition, whereas all the last $k$ bits of it remain
1608: arbitrarily chose-able for the history, provided that $k<r$.
1609: This explains the result $2^{(m_l+m_r)}\cdot 2^k$ for the
1610: number of alternative histories satisfying the shift constraint.
1611: For the entropy of the approximately
1612: decoherent set of histories $\{ h_{\vec{\ys}^1,\,\vec{\ys}^2}\}$
1613: we get the result:
1614: \begin{itemize}
1615: \item Entropy after $k$ iteration steps in case $k\ge m_l+m_r$:
1616: \begin{eqnarray}
1617: H[\{ h_{\vec{\ys}^1,\,\vec{\ys}^2}\}]
1618: &=& -\sum_{\vec{\ys}^1,\,\vec{\ys}^2 }p[ h_{\vec{\ys}^1,\,\vec{\ys}^2}]
1619: \log_2p[h_{\vec{\ys}^1,\,\vec{\ys}^2}] \nonumber \\
1620: &=&k+\;(m_l+m_r)\;+\;{\cal
1621: O}(\frac{(l+r-k)\log_2(l+r-k)}{2^{l-2(k^2+(1+m_l+m_r)k)}})\;.
1622: \end{eqnarray}
1623: \end{itemize}
1624:
1625:
1626: \subsubsection{Hierarchical (multi-scale) coarse-grainings}
1627: \label{sec:Derivations_Hierarchical_coarse-grainings}
1628:
1629: We will see in the following that by introducing more and more
1630: scales that are coarse-grained over in the symbolic representation of
1631: the dynamics the short-term behaviour of the coarse-grained
1632: evolution of the quantum baker's map will exhibit a growing
1633: entropy increase per iteration step, i.e., growing unpredictability.
1634:
1635: So let us now look at the generalised type of
1636: histories~(\ref{set:hierarchically-coarse-grained-histories}).
1637: The evaluation of the corresponding decoherence
1638: functional~(\ref{hierarchical-dfunc}) is done in a similar
1639: way as for the case $\lambda=2$. We first state the result for
1640: the short-term regime which we now define to be given by
1641: $k<\mbox{min}\{m_1,m_2, \dots, m_{\lambda-1}\}$:
1642: \begin{itemize}
1643: \item Decoherence functional in the case
1644: $k<\mbox{min}\{m_1,m_2, \dots, m_{\lambda-1}\}$:
1645: \begin{eqnarray}\label{result_[k<min_m_j]}
1646: {\mathcal D}_{B,\,\rho_0}
1647: [h_{\vec{\ys}^1,\,\vec{\ys}^2,\dots,\vec{\ys}^{\lambda}}\,,\,
1648: h_{\vec{\zs}^1,\,\vec{\zs}^2,\dots,\vec{\zs}^{\lambda}}]
1649: &=&
1650: 2^{-\lambda k}
1651: \cdot\underbrace{\left(\,\prod_{j=1}^{k}
1652: \prod_{i=1}^{\lambda}
1653: \delta_{\ys^{j,i}}^{\zs^{j,i}}
1654: \right)}_{{\rm diagonal}}\cdot
1655: \underbrace{\left(\prod_{i=1}^{\lambda}
1656: \delta{}_{\ys^{1,i}_{1:s_i-1}}^{\xs^i_{2:s_i}}
1657: \right)}_{{\rm first \; shift}}\times
1658: \nonumber\\
1659: && \times
1660: \underbrace{\left(\,\prod_{j=1}^{k-1}\prod_{i=1}^{\lambda}
1661: \delta{}_{\ys^{j+1,i}_{1:s_i-1}}^{\ys^{j,i}_{2:s_i}}
1662: \right)}_{{\rm step-by-step\ shift}}\cdot
1663: \underbrace{\left(\prod_{i=1}^{\lambda}
1664: \delta{}_{\ys^{k,i}_{1:s_i-k}}^{\xs^i_{k+1:s_i}}
1665: \right)}_{k{\rm -th\ shift }}\\&&
1666: \;\; +\,\, {\cal
1667: O}\Big(\frac{l+r-k}{2^{l-2(k^2+(1+m_1+m_2+\dots+m_{\lambda-1})k)}}\Big)\;.
1668: \nonumber
1669: \end{eqnarray}
1670: \end{itemize}
1671:
1672: In the limit of large $l$ the off-diagonal elements of the
1673: decoherence functional vanish and the set of histories becomes
1674: decoherent. The diagonal elements of the functional may therefore
1675: be interpreted as probabilities. The coarse-grained evolution is
1676: again governed by shift constraints. Only such histories are allowed
1677: to arise with significant probabilities that satisfy the shift
1678: condition, which has been illustrated in detail for the case
1679: $\lambda=2$ above. Here we are mainly interested
1680: in the rate of the entropy increase. The result
1681: (\ref{result_[k<min_m_j]}) shows that in the short-term regime,
1682: i.e.\ as long as $k<\mbox{min}\{m_1,m_2, \dots, m_{\lambda-1}\}$,
1683: the coarse-grained evolution exhibits
1684: an entropy increase of $\lambda$ bits per iteration step,
1685: provided that $l$ is very large (classical limit).
1686: This is quantitatively expressed by the entropy of the
1687: approximately decoherent set of histories:
1688:
1689: \begin{itemize}
1690: \item Entropy after $k$ iteration steps in case
1691: $k<\mbox{min}\{m_1,m_2, \dots, m_{\lambda-1}\}$:
1692: \begin{eqnarray}
1693: H[\{
1694: h_{\vec{\ys}^1,\,\vec{\ys}^2,\dots,\vec{\ys}^{\lambda}}\}]
1695: & = & -\sum_{\vec{\ys}^1,\,\vec{\ys}^2,\dots,\vec{\ys}^{\lambda}}
1696: p[ h_{\vec{\ys}^1,\,\vec{\ys}^2,\dots,\vec{\ys}^{\lambda}}]
1697: \log_2
1698: p[ h_{\vec{\ys}^1,\,\vec{\ys}^2,\dots,\vec{\ys}^{\lambda}}]
1699: \nonumber\\
1700: &=&
1701: \lambda\cdot k\;+\;{\cal
1702: O}(\frac{(l+r-k)\log_2(l+r-k)}{2^{l-2(k^2+(1+m_1+m_2+\dots+m_{\lambda-1})k)}})\;,
1703: \end{eqnarray}
1704: \end{itemize}
1705: where we used $p[ h_{\vec{\ys}^1,\,\vec{\ys}^2,\dots,\vec{\ys}^{\lambda}}]
1706: ={\mathcal D}_{B,\,\rho_0}
1707: [h_{\vec{\ys}^1,\,\vec{\ys}^2,\dots,\vec{\ys}^{\lambda}}\,,\,
1708: h_{\vec{\ys}^1,\,\vec{\ys}^2,\dots,\vec{\ys}^{\lambda}}]$.
1709:
1710: For the long-term regime, which we
1711: define by $k>\mbox{max}\{m_1, \dots,m_{\lambda-1}\}$,
1712: our analysis yields the following results:
1713: \begin{itemize}
1714: \item Decoherence functional in case
1715: $k>\mbox{max}\{m_1,m_2, \dots, m_{\lambda-1}\}$:
1716: \begin{equation*}\hspace*{-9cm}
1717: {\mathcal D}_{B,\,\rho_0}
1718: [h_{\vec{\ys}^1,\,\vec{\ys}^2,\dots,\vec{\ys}^{\lambda}}\,,\,
1719: h_{\vec{\zs}^1,\,\vec{\zs}^2,\dots,\vec{\zs}^{\lambda}}]
1720: =
1721: \end{equation*}
1722: \begin{eqnarray}\label{result_[k>min_m_j]}
1723: &=&
1724: 2^{-k-(m_1+\dots+m_{\lambda-1})}
1725: \cdot\underbrace{\left(\,\prod_{j=1}^{k}
1726: \prod_{i=1}^{\lambda}
1727: \delta_{\ys^{j,i}}^{\zs^{j,i}}
1728: \right)}_{{\rm diagonal}}\cdot
1729: \underbrace{\left(\prod_{i=1}^{\lambda}
1730: \delta{}_{\ys^{1,i}_{1:s_i-1}}^{\xs^i_{2:s_i}}
1731: \right)}_{{\rm first \; shift}}\times
1732: \nonumber\\
1733: &&\quad\quad\quad\quad\quad\quad\quad \times
1734: \underbrace{\left(\,\prod_{j=1}^{k-1}\prod_{i=1}^{\lambda}
1735: \delta{}_{\ys^{j+1,i}_{1:s_i-1}}^{\ys^{j,i}_{2:s_i}}
1736: \right)}_{{\rm step-by-step\ shift}}\cdot
1737: \underbrace{\left(\prod_{i=1}^{\lambda}
1738: \delta{}_{\ys^{k,i}_{1:s_i-k}}^{\xs^i_{k+1:s_i}}
1739: \right)}_{k{\rm -th\ shift }}\\&&\quad\quad\quad\quad\quad\quad
1740: \;\; +\,\, {\cal
1741: O}\Big(\frac{l+r-k}{2^{l-2(k^2+(1+m_1+m_2+\dots+m_{\lambda-1})k)}}\Big)\;
1742: \nonumber
1743: \end{eqnarray}
1744: \item Entropy after $k$ iteration steps in case
1745: $k>\mbox{max}\{m_1,m_2, \dots, m_{\lambda-1}\}$:
1746: \begin{equation}
1747: H[\{
1748: h_{\vec{\ys}^1,\,\vec{\ys}^2,\dots,\vec{\ys}^{\lambda}}\}] =
1749: k+\sum_{i=1}^{\lambda-1}m_i\;+\;{\cal
1750: O}(\frac{(l+r-k)\log_2(l+r-k)}{2^{l-2(k^2+(1+m_1+m_2+\dots+m_{\lambda-1})k)}})\;.
1751: \end{equation}
1752: \end{itemize}
1753: The interpretation of these results is similar to the special case $\lambda=2$
1754: of the last section. Whereas in the short-term regime, $k<\mbox{min}\{m_1,
1755: \dots, m_{\lambda-1}\}$, the entropy production rate is $\lambda$ bits per
1756: iteration, in the long-term regime $k>\mbox{max}\{m_1,\dots,m_{\lambda-1}\}$,
1757: the entropy production drops to $1$ bit per iteration, independently of the
1758: values of the parameters $m_1,\dots,m_{\lambda-1}$, which determine the border
1759: between the regimes. In the intermediate regime, the entropy production rate
1760: decreases each time $k$ exceeds one of the values $m_1,\dots,m_{\lambda-1}$.
1761:
1762: \section*{Acknowledgments}
1763:
1764: We would like to thank Todd Brun for helpful
1765: discussions.
1766: This work was supported in part by the European Union IST-FET project EDIQIP.
1767:
1768:
1769:
1770: \begin{thebibliography}{10}
1771:
1772:
1773: \bibitem{Gell-MannHartle1993}
1774: M. Gell-Mann and J. B. Hartle, Phys.\ Rev.\ D {\bf 47}, 3345 (1993).
1775:
1776: \bibitem{BrunHartle1999-PRD}
1777: T.~A. Brun and J.~B. Hartle, Phys.\ Rev.\ D {\bf 60}, 123503 (1999).
1778:
1779: \bibitem{Griffiths1984}
1780: R. Griffiths, J. Stat.\ Phys.\ {\bf 36}, 219 (1984).
1781:
1782: \bibitem{Omnes1988}
1783: R. Omn\`{e}s, J. Stat.\ Phys.\ {\bf 53}, 893, 933, 957 (1988).
1784:
1785: \bibitem{Gell-MannHartle1990}
1786: M. Gell-Mann and J.~B. Hartle, in {\em Complexity, Entropy, and the Physics of
1787: Information}, edited by W.~H. Zurek (Addison Wesley, Redwood City, CA, 1990).
1788:
1789: \bibitem{DowkerHalliwell1992}
1790: H.~F. Dowker and J.~J. Halliwell, Phys.\ Rev.\ D {\bf 46}, 1580 (1992).
1791:
1792: \bibitem{Balazs1989}
1793: N.~L. Balazs and A. Voros, Ann.\ Phys.\ {\bf 190}, 1 (1989).
1794:
1795: \bibitem{Saraceno1990}
1796: M. Saraceno, Ann.\ Phys.\ {\bf 199}, 37 (1990).
1797:
1798: \bibitem{Arnold1968}
1799: V.~I. Arnold and A. Avez, {\em Ergodic Problems of Classical Mechanics}
1800: (Benjamin, New York, 1968).
1801:
1802: \bibitem{Berry1979}
1803: M.~V. Berry, N.~L. Balazs, M. Tabor, and A. Voros, Ann.\ Phys.\ {\bf 122}, 26
1804: (1979).
1805:
1806: \bibitem{Weyl1950}
1807: H. Weyl, {\em The Theory of Groups and Quantum Mechanics} (Dover, New York,
1808: 1950).
1809:
1810: \bibitem{Schack2000a}
1811: R. Schack and C.~M. Caves, Applicable Algebra in Engineering, Communication and
1812: Computing (AAECC) {\bf 10}, 305 (2000).
1813:
1814: \bibitem{Alekseev1981}
1815: V.~M. Alekseev and M.~V. Yakobson, Phys.\ Reports {\bf 75}, 287
1816: (1981).
1817:
1818: \bibitem{Soklakov2000a}
1819: A.~N. Soklakov and R. Schack, Phys.\ Rev.\ E {\bf 61}, 5108 (2000).
1820:
1821: \bibitem{Hartle1998}
1822: J.~B. Hartle, Physica Scripta {\bf T76}, 67 (1998).
1823:
1824: \bibitem{BrunHartle1999-PRE}
1825: T.~A. Brun and J.~B. Hartle, Phys.\ Rev.\ E {\bf 59}, 6370 (1999).
1826:
1827: \bibitem{Brun1995}
1828: T.~A. Brun, Phys.\ Lett.\ A {\bf 206}, 167 (1995).
1829:
1830: \bibitem{Scherer2004b}
1831: A.\ Scherer and A.~N.\ Soklakov, J.~Math.\ Phys.~46, 042108 (2005).
1832:
1833: \bibitem{Soklakov2002}
1834: A.~N.\ Soklakov and R.\ Schack, Phys.\ Rev.\ E {\bf 66}, 036212 (2002).
1835:
1836: \end{thebibliography}
1837:
1838:
1839: \end{document}
1840:
1841:
1842:
1843: