cond-mat0112439/qsm
1: 
2: %%%%%%%  Version 2: A Reference added. 
3: 
4: 
5: %\renewcommand{\thesection}{\Alph{section}}
6: %\renewcommand{\thesubsection}{{\subsection}}
7: \documentstyle[11pt,epsfig]{article}
8: \renewcommand{\theequation}{\thesubsection.\arabic{equation}}
9: \begin{document}
10: \newcommand{\lan}{\langle}
11: \newcommand{\ran}{\rangle}
12: \newcommand{\be}{\begin{equation}}
13: \newcommand{\br}{\begin{eqnarray}}
14: \newcommand{\ee}{\end{equation}}
15: \newcommand{\er}{\end{eqnarray}}
16: \title{
17: \hfill\parbox{4cm}{\normalsize IMSC/2001/12/55\\
18:                                cond-mat/0112439}\\       
19: \vspace{2cm}
20: On the Emergence of the Microcanonical Description
21: from a Pure State}
22: \author{N. D. Hari Dass, S. Kalyana Rama and B. Sathiapalan\\
23:  {\em  Institute of Mathematical Sciences}\\{\em Taramani }\\
24: {\em Chennai, India 600113}}                                    
25: \maketitle                                                                 
26: \begin{abstract}                                                           
27: We study, in general terms, the process by which a pure state can
28: ``self-thermalize'' and {\em appear} to be described by a
29: microcanonical density matrix.  This requires a quantum mechanical
30: version of the Gibbsian coarse graining that conceptually underlies
31: classical statistical mechanics. We introduce some extra degrees of
32: freedom that are necessary for this. Interaction between these degrees
33: and the system can be understood as a process of resonant absorption
34: and emission of ``soft quanta''. This intuitive picture allows one to
35: state a criterion for when self thermalization occurs. This paradigm
36: also provides a method for calculating the thermalization rate using
37: the usual formalism of atomic physics for calculating decay rates. We
38: contrast our prescription for coarse graining, which is somewhat
39: dynamical, with the earlier approaches that are intrinsically
40: kinematical.  An important motivation for this study is the black hole
41: information paradox.
42: \end{abstract}
43: 
44: \section{Introduction}
45: \subsection{Motivation}
46:   Understanding the principles of Statistical Mechanics from first
47: principles still remains a challenge [\cite{reichl}-\cite{Haake}] despite the monumental work by the
48: pioneers like Boltzman, Poincare etc.
49:   According to the fundamental principles of statistical mechanics
50: a macro-system starting from an arbitrary
51: initial state  eventually tends to a state of so called
52: {\it thermodynamic
53: equilibrium}.
54: Inherent in this assertion is a certain 
55: {\em irreversibility}
56: and one of the main issues has been the reconciliation
57: of this with the fact that basic dynamics is 
58: {\em reversible}.
59: In the case of Quantum Statistical Mechanics there are
60: additional complications
61: as the thermodynamic equilibrium state, which
62:  is described by a {\it mixed density matrix}
63: can never be obtained from a  {\it pure}
64: density matrix under quantum mechanical evolution which
65: is described by unitary transformations.
66: This last issue is also the crux of the so called
67: {\em Quantum Measurement Problem}.
68: Interest in these issues has recently been rekindled
69: from a most unexpected direction i.e the problems
70: of black hole entropy and of the so called black hole
71: information. In this section we briefly describe each
72: of these to bring into focus the issues we address in the rest of the paper
73: 
74: 
75: \subsection{Foundations of Quantum Statistical Mechanics}
76: 
77: The questions discussed above lead naturally to questions about the
78: basic postulate of quantum
79: statistical mechanics(QSM). The postulate is that of equal a priori probabilities for all
80: microstates in a given macrostate of definite energy. Equivalently,
81: the density matrix
82: of an ensemble whose energy lies in the range $E_0$ to $E_0+\Delta E$ is given
83: by the identity
84: matrix in the energy basis, suitably normalized as shown
85: ($N$ is the number of states in the given energy interval):
86: 
87: \be   \label{1.1}
88: \rho = {1\over N}I  \; \; \; E_0 \le E \le E_0 + \Delta E
89: \ee
90: and for $E$ not in this interval,
91: \[
92: \rho =0
93: \]
94: It is worth pointing out here that the standard
95: derivation of the microcanonical distribution 
96: can not be carried over automatically to QSM.
97: In classical statistical mechanics one derives the
98: microcanonical distribution, by maximizing the entropy:
99: \be
100: S = -k_B\int dX^N \rho(X^N) log \rho(X^N)
101: \ee
102: subject to
103: \be
104: 1= \int dX^N \rho(X^N)
105: \ee
106: In these eqns $\rho(X^N)$ is the {\em distribution 
107: function} on the {\em phase space}.
108: The result one obtains is
109: \be \rho(X^N) = const ~~~~~ E_0\le H\le E_0+\Delta E
110: \ee
111: and zero otherwise.
112: It is often stated that  for quantum systems the proof
113: is similar with the density operator replacing
114: $\rho$, the quantum mechanical density matrix
115: operator \cite{reichl}:
116: \be
117: S = -k_B tr \rho \log\rho~=-k_B\sum_n P_n\log P_n
118: \ee
119: where $P_n = \lan E,n|\rho|E,n\ran$(the fact that $\rho$
120: is a constant of motion has already
121: been used).
122: %It is stated that
123: %maximizing $S$ wrt $P_n$ yields $P_n = 1/N(E)$.
124: But for a pure state the eigenvalues of $\rho$ are $0,1$
125: (only one eigenvalue 1
126: and all
127: others 0). Hence the entropy is exactly {\it zero} 
128: and no variational principle is applicable.\\
129: 
130: It is sometimes argued that what one deals in practice
131: is not $\rho$ itself but the time-averaged $\rho$:
132: \be
133: \bar\rho(t) = {1\over 2T}\int_{t-T}^{t+T}~dt^{'} ~\rho(t^{'})
134: \ee
135: It follows trivially that
136: \be
137:  tr~\bar\rho = 1
138: \ee
139: But $tr {\bar\rho}^2\ne 1$  as can be easily shown using:
140:  $\rho(t) = e^{iHt}~~\rho(0)~~e^{-iHt}$:
141: \be
142: {\bar\rho(t)}^2 = {1\over 4T^2}\int~dt^{'}~dt^{''}~\rho(t^{'})~ \rho(t^{''})
143: \ee
144: Though $tr~{\bar\rho}^2~\ne 1$ and therefore the time 
145: averaged ${\bar\rho}$ represents a
146: mixed state,
147: $tr~{\bar\rho}^2$ is calculable in terms of the initial density matrix:
148: \be
149: tr~~{\bar\rho(t)}^2~~= \sum_{n,m}~~|\lan n|\rho(0)|m\ran |^2({\sin(\Delta E_{nm} ~~T)\over
150: (\Delta E_{nm}~~T)})^2
151: \ee
152: In fact traces of all powers of ${\bar\rho}$ are calculable
153: and found to be completely determined by the initial
154: pure density matrix and the spectrum of the Hamiltonian
155: describing the system. The important point is that all
156: these constraints have to be used in the variational
157: method and subsequently one does not get the
158: microcanonical distribution.
159: 
160: One can ask, what is the nature of a quantum mechanical
161: system that has the property that, even though we start
162: it off in a pure state, it evolves after some suitable
163: time $\tau$ to a state that {\em for all practical 
164: purposes} (which means for all observables of interest) gives the
165: {\em same expectation} values as a system described 
166: by the
167: microcanonical density matrix.
168: To mimic the situation in classical statistical mechanics
169: where we consider an initial configuration with energy
170: lying in a narrow band,
171: it is assumed that the pure state is a linear combination
172: of energy eigenstates with energy between $E_0$ and
173: $E_0+\Delta E$. We should also discuss what happens when
174: the pure state itself happens to be an energy 
175: eigenstate. We will
176: give an answer stated in terms of the properties of
177: the exact eigenfunctions of the system. 
178: This will be the main result of this paper. 
179: It seems to us that there
180: is an inbuilt dependence on the choice of
181: ``interesting'' observables which define the coarse-grained
182: microstates.
183: \subsection{Quantum Measurement Problem}
184: According to the von Neumann {\it projection postulate}, also
185: known as the '{\it collapse of the wave-function}' 
186: postulate,
187: if an observable $A$ is measured in a generic quantum 
188: state,
189: the result will be any one of the eigenvalues of $A$ 
190: and the
191: state after the measurement {\it collapses} to the 
192: corresponding
193: eigenstate. This implies a transition from the initial
194: {\it pure ensemble} characterized by
195: \be
196:  tr~ \rho = 1, tr~ \rho^2 = 1
197: \ee
198: before the measurement to a {\it mixed ensemble}
199: with
200: \be
201: tr~ \rho = 1, tr~ \rho^2\ne 1
202: \ee
203: Let us illustrate this with a simple example. Consider 
204: an ensemble of states $|\psi>$ of a two-level system 
205: on which a measurement is done whose possible outcomes 
206: are $s=\pm$ with the corresponding eigenstates 
207: $|+>,|->$.
208: As the latter span a basis for the two-level Hilbert 
209: space
210: we could expand $|\psi>$ in this basis
211: \be
212: |\psi>~=~c_+|+>~+c_-|->
213: \ee
214: The initial density matrix
215: \be
216: \rho_i=|\psi><\psi|
217: \ee
218: is {\em pure}.
219: After the measurement we have a {\it mixture} of two 
220: pure ensembles of states $|+>,|->$ with weight 
221: factors $|c_+|^2, |c_-|^2$ resulting in the 
222: final density matrix
223: \be
224: \rho_f= |c_+|^2~|+><+|~+|c_-|^2~|-><-|
225: \ee
226: It is easy to check that
227: \be
228: tr~ \rho_f = 1, tr~ \rho_f^2=(|c_+|^2)^2+(|c_-|^2)^2
229: = 1-2|c_+|^2|c_-|^2\leq 1
230: \ee
231: 
232: As  {Unitary} transformations preserve all traces
233: \be
234: \rho^{'} = U\rho U^{\dagger}\rightarrow tr~ {\rho^{'}}^n
235: = tr~ {\rho}^n
236: \ee
237: the measurement process is not describable by a unitary
238: transformation and is thus {\it irreversible}. It is as
239: if the measurement process has to be treated 
240: differently from ordinary dynamics. This is the crux 
241: of the so called {\em Quantum Measurement Problem} 
242: and one sees that once again the crucial
243: issue has arisen of pure states evolving into 
244: mixed states,
245: or more plausibly into pure states that {\em for all 
246: practical purposes} look like mixed states.
247: \subsection{Black hole evolution and AdS/CFT 
248: correspondence}
249: One of our strong motivation for this study comes 
250: from an issue related to
251:  the ``information paradox'' in black hole physics.
252: The issue is the following: Consider the quantum 
253: mechanical
254: description of black hole
255: formation. Some matter/energy in some initial state
256: described by a wave function evolves in time and at 
257: some point
258: makes a transition to a black hole state due to 
259: the attractive gravitational interaction. As per the
260: Hawking effect the black hole 
261: appears to be radiating
262: like a blackbody at its Hawking temperature. The 
263: system does not look to an external observer to be 
264: described by a pure state wave function but rather by 
265: a mixed state density matrix characterstic of a thermal
266: state. The usual argument is that this transition
267: from a pure state to a mixed state is {\it illusory} 
268: because when we include the degrees of freedom 
269: {\it inside} the black hole one recovers a pure state
270: description. Indeed string theory has given us a 
271: prescription in some situations for actually
272: counting the number
273: of microscopic states associated with a black hole 
274: which reproduces the Bekenstein entropy. Nevertheless,
275: (and this is the crucial point), while ignoring the 
276: degrees of freedom in the
277: interior of the
278: black hole can make the pure state appear mixed to 
279: the external observer,
280: this does not explain
281: why it should look {\it thermal}. In other words, 
282: when calculating the entropy
283: by counting the
284: number of states, there is an {\em implicit} 
285: assumption that the interior system
286: is a {\em mixture} of states with {\em equal} a priori
287: probabilities i.e it is ergodic and is described by a 
288: microcanonical ensemble. In
289: ordinary statistical systems
290: there is always a ``heat bath'' one usually invokes - 
291: basically the environment - that will ensure
292: this, but in the case of the black
293: hole we are describing a closed system. There is no 
294: environment or heat bath.
295: So the question is
296: how does one justify such an assumption of ergodicity 
297: for a closed system?
298: 
299: In \cite{KRBS} this problem was approached using the 
300: so called AdS/CFT correspondence \cite{maldacena,witten}. Using this
301: correspondence the gravitational problem is mapped 
302: to a Yang-Mills problem. How does a pure state with
303: some fixed energy in a Yang-Mills theory 
304: ``self-thermalize''? The solution proposed
305: was that chaos
306: would do the job. Classical Yang-Mills has already been
307: shown to be chaotic
308: \cite{Biro,Muel}. One
309: expects chaos
310: to develop and make the system ergodic. Assuming this 
311: is true quantum mechanically also,
312: this phenomenon could be mapped back to the black-hole via the same AdS/CFT correspondence.
313: 
314: While not much is known theoretically about quantum 
315: chaos in Yang-Mills theories there is
316: some experimental
317: evidence in heavy ion collisions for the formation of 
318: a quark-gluon plasma (at finite temperature).  The 
319: initial state, which is two heavy nuclei traveling
320: towards each other, is definitely described by a 
321: quantum mechanical wave function.
322: The
323: final state appears to be describable by a system
324: at finite temperature. If this happens, then it is 
325: quantum mechanical
326: ``self thermalization'' - a pure state evolves 
327: unitarily by Hamiltonian evolution
328: and after
329: a while {\em looks like} a thermal state.
330: 
331: 
332: \subsection{Criterion for a Physical System to Become 
333: Ergodic.}
334: 
335: Given that we are able to characterize the nature of 
336: eigenfunctions for a system that
337: is ergodic, we can ask when does a physical system satisfy these properties ?
338: i.e. when
339: do we expect thermalization? 
340: We will give an approximate answer to this question
341:  in this paper.
342: The centre of mass degrees of freedom are expected to look thermal
343: most of the times. The interesting question concerns the {\em internal}
344: degrees of freedom that are normally frozen to some discrete quantum states.
345: In the Yang-Mills example above, we expect
346: thermalization of quark and gluon degrees of freedom to
347: take place only when the energy density exceeds a critical density. The scale is
348: obviously set by $\Lambda _{QCD}$. Similarly in the black hole example
349:  we do not expect
350: every (zero temperature) neutron star to form a black hole and produce a non-zero
351: Hawking temperature. This is discussed further in sec. 2.6.4.
352: 
353: \subsection{Thermalization Time}
354: 
355: If a physical system is expected to become ergodic, 
356: we can ask what is the time scale over which this 
357: happens. This is also the time scale for return to 
358: equilibrium when the system is disturbed. It should of
359: course be emphasised that depending on how the system
360: is disturbed, there could be many such time scales. We 
361: will give approximate answers to such questions.
362: 
363: \subsection{Non-Quantum-Mechanical and 
364: Quantum-Mechanical Coarse Graining}
365: 
366: There have been several attempts to obtain effective 
367: thermal density matrices, or, to obtain irreversibility
368: from reversibility. All these involve (as they must 
369: indeed) some coarse graining. However, to our knowledge
370: many of these involve ad hoc prescriptions that do not
371: arise {\em naturally} in quantum theory.  For example, 
372: some invoke averaging over the time of measurement. The
373: argument is that every measurement takes a finite
374: time. Nevertheless averaging over time alone is not 
375: sufficient.
376: As described in Section 1.2 it also does not give
377: the right answer.
378: Similarly, some invoke
379: averaging over an ensemble of initial states. 
380: In quantum mechanics there is no need for any other
381: ensemble than that required by the probabilistic
382: interpretation of the quantum state. Thus averaging
383: over some distribution of initial states is generically
384: ad hoc and unwarranted. The point is that in 
385: generic situations there is no justification for such 
386: ad hoc averaging procedures.
387: Thus whatever
388: coarse graining is necessary must arise naturally 
389: within the framework of quantum mechanics.
390: We will see that invoking some unobserved
391: ``soft'' (low energy) degrees of freedom can naturally
392: accomplish the required coarse graining. An example 
393: of such degrees of freedom is
394: the ``soft'' photons of QED.
395: 
396: \subsection{Outline}
397: In Section 2 we will describe our proposal for Gibbsian
398: coarse graining in
399: quantum mechanics using
400: soft quanta. This will
401: address the criticisms of section 1.7. We will also 
402: give a physical
403: criterion for thermalisation-
404: a qualitative answer to the question of Section 1.5.
405: In Section 3 we introduce the mathematical formulation 
406: of the problem.
407: We also give a brief description of the work of 
408: von Neumann
409: and van Kampen on this subject (about which we learnt 
410: well after completing our work ). This is included 
411: mainly for completeness and comparison,
412: and is not logically required for understanding 
413: our work. In Section 4
414: we  discuss the
415: issues raised in Section 1.2 regarding the 
416: characteristics of a system that
417: can be described by quantum statistical mechanics.  
418: In Section 5
419: we look at the same question from a different 
420: perspective - that of
421: soft quanta and resonant transitions.
422: In Section 6 we will discuss a two level system
423: coupled to a continuum of soft quanta states. This will illustrate some of the ideas
424: more quantitatively. By using the quantum mechanical 
425: formalism underlying the familiar ``Fermi Golden
426: Rule'', we will see approximate irreversibility coming 
427: out of a reversible dynamics
428: and the approximate emergence of a thermal 
429: microcanonical density matrix.
430: More quantitative answers to the questions of 
431: Section 1.5,1.6 and
432: 1.7 will be given here.
433: Section 7 will summarize the results of this paper.
434: 
435: \section{Gibbsian Coarse Graining and the Importance of Soft Modes}
436: \setcounter{equation}{00}
437: 
438: \subsection{Classical Gibbsian Coarse Graining}
439: 
440: An important ingredient in classical statistical mechanics is the notion
441: of coarse graining. As was originally argued by 
442: Gibbs \cite{gibbs}, unless the microstates
443: are coarse grained, entropy will always remain 
444: constant. This is easy to see
445: - as the region of phase space( energy surface) 
446: that is occupied by the ensemble (call this
447: region $\Gamma^*$; care should be taken to distinguish
448: $\Gamma^*$ from $\Gamma$ which usually denotes the
449: {\em entire} energy surface) spreads
450: out in phase space, its shape becomes very complicated 
451: but its
452: volume remains fixed. 
453: The  entropy  is given by $S= \sum _i p_i ln \; p_i$ 
454: where $i$ labels the microstate
455: and $p_i$ is the probability that the system is 
456: in this state.
457: If microstates are taken to be 
458: points, then 
459: $p_i=1$ if the point $i$ belongs to $\Gamma^*$ and zero
460: otherwise. It is easy to see that this number is 
461: constant because it depends only on the
462: volume of $\Gamma^*$, not the shape. If, on the other 
463: hand, the microstates are taken to be
464: small boxes(but large enough to have many system points
465: within them) in phase space, then $p_i$ will be equal 
466: to the fraction of the
467: box that is inside $\Gamma^*$. In this case, it is clear that as $\Gamma^*$  
468: spreads, most of the $p_i$'s will become equal to each other and approach
469: a value between 0 and 1. Thus $S$ will increase till it reaches a maximum.
470: 
471: \subsection{Quantum Mechanical Gibbsian Coarse Graining- ``Soft Quanta''}
472: 
473: What we need is the quantum mechanical analogue  of this coarse graining.
474: It is  important that one should stay within the formalism of quantum
475: mechanics while describing this coarse graining. We propose the following scheme:
476: Let us consider our Hilbert space to be a {\em tensor}
477: product of two Hilbert spaces ${\cal H}_i$ and 
478: ${\cal H}_a$ i.e ${\cal H}={\cal H}_i\otimes{\cal H}_a$.
479: States in ${\cal H}$ are the {\em microstates} of
480: our system. $i$ are the degrees of freedom that one is 
481: physically
482: interested in and will represent the "coarse grained"
483: microstates. $a$ represents some degrees of freedom that we are not
484: interested in and possibly over which we have no control.
485: These are the degrees that allow us to coarse grain. As an example
486: consider a gas of molecules. $i$ could be the usual microscopic degrees
487: of freedom that one
488: associates with the gas, the positions and momenta of the molecules. One
489: can also include rotational or vibrational degrees if one wants. $a$
490: can be soft photons. i.e. the molecules interact with the electro-magnetic
491: field all the time and constantly emit and absorb radiation. There are
492: very long wavelength photons of almost zero energy that one has no control
493: over. They constitute a continuum of gap-less 
494: excitations in this system.
495: They interact with the molecules but take away negligible energy. The energy
496: hyper-surface defining the microcanonical ensemble in quantum mechanics
497: always has a small but finite width.
498: We can understand this as follows: We can specify the variables $i$.
499: The $a$ variables are not in our control. We can think of $i$ as labeling
500: an energy eigenstate of the $i$ system when there is no interaction with the
501: $a$ system.  When we turn on interactions with $a$, the state $i$ is no longer
502: an energy eigenstate. It becomes a linear combination of states with energy
503: in a range
504: $\Delta E$. This is an estimate of $\Delta E$ in (\ref{1.1}).
505: The observables of interest $O$ will be assumed to be
506: functions only of $i$ and not of $a$. To be precise we assume that
507: \be \label{CG}
508: \langle i,a | O | j,b \rangle = \delta _{ab}\langle i |O|j \rangle
509: \ee
510: 
511: Thus our coarse graining will be defined by saying that these are the
512: operators of interest and it is with respect to these operators that
513:  the system looks thermal. Note that this is different from the following
514: kind of coarse graining: the observable $O$ depends on both $i$ and $a$, i.e.,
515:  $\lan  i,a| O | j,b \ran $ is not necessarily given by \ref{CG}.
516:  Nevertheless, the $a$ variables are not in our control. So we average
517:  over them in some fashion - for e.g. $ {\bar O _{ij}} = \int da db
518:  \lan i,a | O | j,b \ran p_{a,b}$. This defines a coarse graining. But
519:  this is ad hoc and definitely is not
520:  a quantum mechanical operation. We do not want this type of coarse graining.
521: %---
522: 
523: There is some arbitrariness in what we call $i$ and what we call $a$. But
524: it is crucial that the $a$ degrees are gap-less, or at least the gap
525: $\delta _a$ between two consecutive energy levels of the
526: unperturbed $a$ system should be much smaller than
527: the inverse of the time interval during which we observe the system.
528: Thus for instance, if $\delta _i =0$ for some of the 
529: $i$-states, then some of the $i$ degrees of freedom
530: could be called $a$.
531: Thus if $i$ represents a continuum of harmonic oscillators, then some subset
532: of these around zero frequency can be called the $a$ variables. Since
533: our energy resolution is always finite, these are not observables of interest.
534: 
535: \subsection{Paradigm of Discrete States Coupled to a Continuum}
536: 
537:  If  it is the case that we can describe the exact system as
538: a coupled $i$-$a$ system with $i$ being discrete and $a$ being continuous, 
539: then
540: we are familiar with this in atomic physics situations. In this situation if
541: we focus on the discrete system alone it will appear to be described by
542: a non-Hermitian Hamiltonian with complex energy eigenvalues. The width
543: $\Gamma$ represents the finite lifetime of the (excited) states. This
544:  requires
545: that we take the limit $\delta _a \rightarrow 0$. If we keep $\delta _a$
546: finite then we expect that after a finite time of O(${1\over \delta_a}$)
547: the system will look periodic. Thus the apparent irreversibility is only
548: because the time of observation is short compared to $1\over \delta _a$.
549: Conversely if we are to observe an irreversibility, as in the second law
550: of thermodynamics, it is clear that
551:  we need such soft degrees of freedom.  ``Soft'' in this context
552: means that the energy of these quanta should be less than
553: the bandwidth, $\Delta E$,
554: that defines the microcanonical system as :
555: $E_0 - \Delta E \leq E \leq E_0 +\Delta E$.
556: Thus we want $\hbar \omega < \Delta E$.
557: In all the
558: usual physical systems there are always soft quanta. So this is not an
559: issue.
560: 
561: \subsection{Heuristic Criterion for Ergodic Density Matrix}
562: 
563: However the above condition is not sufficient. Arguments in the next section
564: suggest that
565: matrix elements between $|i \rangle$ and $|j\rangle$ induced by the
566: interactions with $a$ should be much larger than $\delta _i$. This ensures that
567: an exact energy eigenstate contains a large number of different $i$ states.
568: This is not
569: always satisfied. Systems that do not satisfy this will not be ergodic.
570: In a
571: quantum system with an energy gap (particles in a box) one does not expect
572: any kind of ergodicity when the total energy of the
573: system is such that only the lowest
574: energy levels
575: are excited. \footnote{We remind the reader that we are discussing
576: closed systems and there is no heat bath.}
577: 
578: \subsection{Resonance and Soft Quanta}
579: 
580: When the energy of the soft quanta is equal to $\delta _i$, we
581: expect resonant absorption/emission to take place.  This induces the large
582: off-diagonal matrix elements between $|i\ran $ and $|j\ran$ states that
583: we referred to above.  Because of resonance, the coupling between the soft
584: quanta and the $i$ system need not be large for this to happen.
585: Furthermore we expect
586: both states $|i\ran$ and $|j\ran$ to be equally populated if the
587: probability, $P_{ij}$
588: of $|i\ran \rightarrow |j\ran$ is equal to
589: $P_{ji}$. This will be the case when the induced emission dominates the
590: spontaneous emission, which means that there should be a large number
591: of soft quanta available, i.e. the energy/degree of freedom in the $a$-system
592: should be much larger then the largest of the energy gaps $\delta _i$
593: which is $\Delta E$. 
594: Thus the criterion for thermalization is:
595: 
596: i)there should be a {\em continuum} of soft quanta
597: with energies in the range
598: $\delta _i$ to $\Delta E$. This ensures resonant transitions.
599: 
600: ii)The number of soft quanta in each mode should be $>>1$. This ensures the equality
601: of the upward and downward transitions and hence equality of
602: occupation probability.
603: 
604: 
605: In section 6 we will study
606: a quantum system where some calculations can be done perturbatively. The
607: results support the general picture.
608: 
609: \subsection{Examples:}
610: 
611: 
612: \subsubsection{$H_2$-molecules}
613: 
614: Let us turn now to some physical examples to illustrate these ideas.
615: Consider a gas of hydrogen molecules.
616: Clearly, as far as the centre of mass
617: degrees of freedom of the molecules are concerned the system is 
618: likely to be ergodic. This is again a situation where $\delta _i =0$.
619: When $\delta _i=0$ the off-diagonal elements of the full
620: Hamiltonian are clearly larger than $\delta _i$ and one expects ergodic
621: behaviour by the above arguments. The centre of mass motion
622: is in any case ergodic.
623: 
624: Let us focus our attention on the relative coordinates of the atoms
625: in one molecule and consider this as the $i$ system.
626: We would like to understand whether 
627: ``self thermalization'' can take place for this sub-system.
628: We could alternatively
629: consider all the molecules, but the interaction between the internal
630: degrees of freedom being negligible, this will just be many copies
631: of the one molecule system.  
632:  The
633: $a$ states are a subset of the degrees of freedom
634: of the rest of the gas molecules that interact with the internal degrees
635: of this molecule. We will coarse grain over these degrees
636: by considering observables that  depend only on $i$. We
637: assume that the energy exchange
638: between the internal degrees of this molecule and that of the rest of the gas
639: molecules (which includes the `$a$` degrees) is very small -
640: smaller than the resolution of our
641: experiment. It therefore makes sense to talk of the microcanonical density
642: matrix for the $i$-system.
643: Let us assume that the kinetic energies
644: of the molecules are small and all center of mass
645: degrees can be included in the $a$-system. \footnote{In a
646: realistic system it may be that these conditions are satisfied at
647: such low ``temperatures'' that the $H_2$-gas may have liquefied. But these
648: considerations can apply for any state of matter. }
649: 
650: We assume that the potential energy and the energy 
651: levels of this system
652: are as shown in figure 1. The width of the dashed line
653: represents the energy resolution $\Delta E$ that defines a microcanonical
654: system.  The energy of the soft `$a$` quanta has to be less than $\Delta E$
655: 
656: 
657: \begin{figure}[htbp]
658: \begin{center}
659: \epsfig{file=qsmfig1.eps, width= 12 cm,angle=0}
660: \vspace{ .2 in }
661: \begin{caption}
662: {In the energy range ``a'' the system is unlikely to be
663: ergodic. In ``c'' it is very likely ergodic,
664: and in ``b'' it is almost ergodic. The width of the dashed
665: line represents $\Delta E$   }
666: \end{caption}
667: \end{center}
668: \label{fig1}
669: \end{figure}
670: 
671: 
672: 
673: If the energy of the system described by the Hamiltonian for relative
674: motion is in the region shown by the dashed line marked ``a'',
675: where $\delta _i$ is fairly large compared to $\Delta E$ we do
676: not expect ergodic behaviour. We expect that if we start the molecule off
677: in a pure state, such as the ground state or first excited state,
678: it will remain there. This follows from purely energetic considerations.
679: 
680: The region marked ``c'' is a continuum
681: where the molecule has dissociated and we have atomic Hydrogen. In this
682: region $\delta _i \approx 0 << \Delta E$ and one expects ergodicity.
683: 
684: 
685: Region marked ``b'' is in between, with $\delta _i \approx \Delta E$
686: and one can expect ergodicity if
687: the off diagonal matrix elements are large compared to $\delta _i$. As long as
688: there are soft quanta that can be absorbed or emitted whose energies
689: match the energy spacings ($\delta _i$), this will be the case. The soft quanta
690: are provided by the inelastic collisions with other molecules.
691: We thus expect that even if we start
692: the system off in a pure state it will soon make transitions to the
693: numerous other states in that energy band and be effectively
694: described by a microcanonical
695: density matrix. This is what we would like to demonstrate in Sections 4,5 and 6.
696:  
697: As the above example illustrates, the centre of mass degrees
698: in a gas, for instance, typically
699: are expected to be ergodic. They correspond to the $\delta _i =0$
700: case. Classical analysis of ``billiard balls'' in certain
701: situation shows ergodicity
702: even with a small number of balls \cite{Sinai}. We need not expect
703: the quantum behaviour of this weakly interacting system to be very different.
704: 
705: 
706: The difficult question  concerns  the non-center-of-mass degrees that have
707: non-zero $\delta _i$. Are they ergodic or not?
708:  It can happen sometimes that
709:  while the center-of-mass degrees are ergodic, the
710:  non-center-of-mass degrees are in a pure state. The point we are
711: trying to make is that the value of the energy spacing $\delta _i$,
712: is a crucial
713: factor in deciding the answer.
714: 
715: 
716: 
717: \subsubsection{QCD Plasma}
718: 
719: Another example is that of heavy ion collision producing quark gluon plasma 
720: Clearly when the energy of collision is smaller than the energy gap
721: in QCD the hadrons are still usefully described by wave functions.
722: Once this gap is exceeded one can expect thermalisation. What are the soft
723: quanta here? Presumably these are soft gluons. There are some points of
724: similarity between this system and the hydrogen molecule example discussed
725: above. Deconfinement here, presumably corresponds to the dissociation of the
726: molecule there.
727: 
728: 
729: 
730: 
731: \subsubsection{Sodium Vapour}
732: 
733: 
734: Let us take another example, where again we expect to see two different
735: kinds of ergodicity. As it turns out this example has 
736: some similarity with the black hole
737: situation to be discussed later. Consider a vapour of sodium atoms. The centre of mass motion
738: of the atoms is ``gap-less'' and can be expected to be ergodic motion,
739: even at low temperatures. The electrons
740: in the atom however are described by pure state wave functions, at temperatures
741: that are low compared to electronic level spacings. Consider what happens when
742: the vapour cools down sufficiently that it forms a Sodium crystal, at some low
743: temperature. The centre of mass motion is no longer ergodic. However sodium
744: being a metal, the outer electrons are free to move around in the crystal.  
745: They
746: are now described by a Fermi gas and their motion is ergodic!
747: One can presumably associate a temperature with them.
748: The gap $\delta _i$ between electronic levels has changed from, say, a
749: few electron volts to something of the order $1\over L$ where $L$ is the
750: macroscopic size of the crystal. This is a quantum mechanical effect. Thus 
751: a phase transition in the material has introduced new gap-less excitations
752: that demonstrate ergodic behaviour, even though at a higher temperature
753: they were not ergodic (because they were confined to the interior of
754: the sodium atom)! 
755: 
756: 
757: \subsubsection{Stars and Black Holes}
758: 
759: 
760: There is a similarity between the sodium vapour example and the black hole example.
761: When matter collapses to form a star the centre of mass motion of the
762: lumps of matter once again can be assumed
763: to be ergodic and as a consequence we get a hot star.
764: (Indeed astrophysicists routinely treat this motion entirely classically
765: using classical statistical mechanics, until we get to high densities.)
766: At some point the star (if sufficiently massive)
767: cools to essentially zero temperature, undergoes
768: a phase transition and forms a black hole. At this point some other
769: internal degrees of freedom (D-branes in string theory black holes)
770:  become (presumably gap-less and) ergodic
771: and there is a temperature that can be associated with them. These are
772: like the outer electrons of the sodium atom. 
773: 
774: 
775: \subsection{Summary}
776: 
777: Let us summarize the points made in this section:
778: 
779: 1. Coarse graining is done by introducing soft quanta $a$ in addition to
780: the conventional microscopic degrees $i$. These soft modes induce transitions
781: between the $|i\rangle$. The gap $\delta _a $ must be very small compared
782: to the inverse time of observation. Also
783: the off diagonal terms in the Hamiltonian connecting different $i$ states
784: must be much larger
785: than the unperturbed gap $\delta _i$. This is necessary for the exact energy
786: eigenstates to satisfy the properties assumed in Section 3.
787: This is thus necessary for ergodicity. In the resonance picture this will automatically
788: be the case if there is a sufficient number of soft quanta with a continuum
789: of frequencies in the
790: appropriate energy range.
791: 
792: 2. The interaction with the continuum $a$ degrees of freedom,
793: induces an anti-Hermitean part to the
794: effective Hamiltonian of the $i$ system. This makes the evolution {\em look}
795: irreversible. The apparent lack of reversibility has
796: to do with treating the $a$-degrees as 
797: {\em continuous}. If a gap $\delta_a$ is introduced,
798: the recurrence time is $O({1\over\delta_a})$.
799: 
800: In section 6 
801: we will calculate, 
802: in a simplified model,
803: the time evolution
804: of the partially traced density matrix that describes the $i$ system and show
805: how it appears to evolve in time to a thermal one.
806: 
807: 
808: \section{Mathematical Formulation of the Problem and Earlier Approaches}
809: \setcounter{equation}{00}
810: 
811: In this section we give a mathematical formulation
812: of the problem we wish to address in this paper. It
813: is helpful to start with a brief recapitulation of the
814: situation in classical statistical mechanics. In
815: particular we shall follow the Gibbsian route formulated
816: on the classical phase space and even there we will
817: be concerned only with the microcanonical ensemble
818: description.
819: 
820: In such a situation one considers an isolated system
821: with its total energy in a narrow band
822: $E_0-\Delta<E<E_0+\Delta$. Under time evolution governed
823: by the dynamical equations the system moves on
824: this hyper-surface of the $6N$-dimensional phase space
825: called the "energy surface". Different regions of this
826: energy surface can be labeled by values of observables
827: other than the Hamiltonian. Generically these
828: observables are time-dependent. Further, the accuracies
829: with which these observables can be determined make it
830: meaningful to decompose the energy surface into
831: elementary cells called "phase cells" and all system
832: points within the same cell are ascribed the same value
833: for the "coarse grained" observables
834: Due to dynamics the system point moves from
835: phase space cell to phase space cell.The crux of the
836: Gibbsian statistical mechanics is the so called
837: ergodic theorem( more precisely the quasi-ergodic
838: theorem) which states that after a sufficiently long
839: time the system point passes through all the phase
840: cells and furthermore the time spent in each phase cell
841: is proportional to its volume. This immediately allows
842: one to equate the time average of observables with
843: an "ensemble average" wherein the weight factor for
844: each phase cell is its volume. As the phase cells
845: can be constructed with equal volume the ensemble
846: average can be taken with a distribution giving
847: equal weights to the different phase cells. This is the
848: microcanonical distribution. The crucial point is that
849: the time average has been replaced by an ensemble
850: average in such a way that the distribution is
851: {\em independent} of the initial state. It should be
852: emphasized that in addition to the quasi-ergodic
853: theorem one also approaches the same problem through
854: the so called H-theorem and the concomitant concept
855: of entropy.In what follows we shall only concentrate
856: on the quasi-ergodic aspects of the problem.
857: 
858: Having stated all this it is very important to
859: emphasize that a precise proof of the quasi-ergodic
860: theorem and in particular the resolution of the precise
861: role of dynamics is a very difficult problem. Both in
862: this formulation as well as in the Boltzmannian approach
863: to statistical mechanics assumptions have to be made
864: that are equivalent to the assumption of chaotic
865: behaviour.
866: \subsection{Formulation of Quantum Statistical Mechanics.}
867: Now the main question is how to formulate and prove
868: similar statements in the context of quantum mechanics.
869: The foremost difficulty here is that unlike as in
870: classical mechanics there is no concept of a phase
871: space nor of a trajectory.If classically one views
872: the phase space as the space of all possible states
873: of the classical system, the natural analog in the
874: quantum case is the Hilbert space. Already at this
875: stage many crucial differences appear; one such is
876: the fact that for even a simple system like the
877: ideal gas while the classical phase space is
878: {\em finite-dimensional} with dimension $6N$, the
879: quantum mechanical Hilbert space ${\cal H}$ is
880: the tensor product of $N$ copies of Hilbert spaces
881: ${\cal H}_i$ each of which is {\em infinite-dimensional}.
882: 
883: Let $H$ be the {\it exact} Hamiltonian of the system and let $|A\ran$
884: be the {\it exact} eigenstate with eigenvalue $E_A$. Now Consider energy
885: eigenstates such that their eigenvalues are in the range
886: $E_0-\Delta<E_A<E_0+\Delta$. Further, let the initial state of the system
887: $|\psi\ran$ be such that
888: \be \label{1}
889: |\psi\ran = \sum_A C_A^\psi |A\ran
890: \ee
891: Then
892: \be \label{2}
893: |\psi(t)\ran = \sum_A~C_A^\psi~e^{-iE_At}~|A\ran
894: \ee
895: If we denote $C_A^\psi(t)=C_A^\psi e^{-iE_At}$ it
896: is obvious that $|C_A(t)|^2$ are {\em independent}
897: of time {\em for all A}. The average energy of the
898: system at any time t is given by
899: \be \label{3}
900: \bar E(t) = \sum_A |C_A(t)|^2 E_A
901: \ee
902: >From eqn (\ref{3}) we see another principal difference
903: between the classical and quantum situations: in the
904: quantum case not only is $\bar E$ constant in time, but
905: so are all the quantities $|C_A|^2$. If we now decompose
906: $C_A(t)$ into its magnitude $r_A$ and its phase $\theta_A$
907:  we find that while $r_A$ does not change with time,
908: $\theta_A(t)=\theta_A(0)-E_At$.
909: 
910: >From this it follows that under quantum dynamics i.e
911: Schr\H{o}edinger equation, the motion of the system point
912: is {\em not over} the entire Hilbert space but over the
913: subspace defined by constant $|C_A(t)|^2$ for each $A$.
914: The motion is in fact on the $N_A$-torus spanned by
915: the angle variables $\theta_A$ where $N_A$ is the number
916: of (non-degenerate)energy eigenvalues in the interval
917: considered. Furthermore, the motion on this $N_A$-torus
918: is such that the angular velocity ${d\theta(t)\over dt}$
919: is constant in every direction and equal to $-E_A$. For
920: a macroscopic system the Hamiltonian $H$ will be
921: generically so complex that this motion will densely
922: fill the entire $N_A$-torus and because of the uniform
923: velocity in every direction the time spent by the system
924: point in any interval $[\{\theta_A\},\{\theta_A+\delta\theta_A\}]$
925:  is exactly proportional to the {\em volume}
926: of the interval. In this sense the quantum motion
927: is quasi-ergodic. The precise conditions to be fulfilled
928: by the spectrum of eigenvalues of $H$ will be discussed
929: later but it suffices to stress here that they are
930: fairly generic and unlike the classical case do not
931: require special assumptions about {\em chaos}.
932: Now consider the expectation value of an observable $O$
933: in the state $|\psi(t)\ran$:
934: \be \label{4}
935: \lan\psi(t)|O|\psi(t)\ran = \sum_{A,B}~C_B^{*\psi}~
936: C_A^ \psi~e^{-i(E_A-E_B)t}~~O_{BA}
937: \ee
938: and a time-average of this expectation value over a
939: duration $T$ centred at time $\tau$
940: \br \label{5}
941: \bar O(\tau) &=& {1\over 2T}~\int_{\tau -T}^{\tau +T}~~
942: \lan\psi(t)|A|\psi(t)\ran\nonumber\\
943: & =& \sum_{B,A} C_B^{*\psi}~C_A~O_{BA}~e^{-i(E_A-E_B)\tau}
944: {\sin {(E_A-E_B)T}\over (E_A-E_B)T}
945: \er
946: Now the main problem is that this expression has an
947: explicit dependence on the parameters $C_A^\psi$ of
948: the initial state and {\em can not} be replaced by an
949: ensemble average that is {\em insensitive} to the {\it
950: initial state}. This is the {\em mathematical 
951: formulation}
952: of the problem to be solved i.e interpret quantum
953: statistical mechanics in such a way that the
954: time-average {\em does not remember} the initial state.
955: 
956: It is necessary to make more precise the notion of
957: the time-average. Note that
958: \be \label{6}
959: {\sin(ET)\over(ET)}_{T\rightarrow\infty}\rightarrow \delta(E)
960: \ee
961: The time average in eqn (\ref{5}) then becomes
962: \be \label{7}
963: \bar O(\tau) = \sum_A |C_A^\psi|^2 O_{AA}.
964: \ee
965: At this stage nothing more can be said. If we however
966: consider some special class of systems it is possible
967: to make more statements.
968: For example, Berry \cite{Berry} has conjectured
969: that for {\em classically chaotic} systems energy
970: eigenfunctions behave like Gaussian random variables.
971: More precisely consider expanding the exact energy
972: eigenstates $|A\ran$ in some {\em orthonormal basis}
973: $|I\ran$
974: \be \label{8}
975: |A\ran = \sum_I C_A^I|I\ran
976: \ee
977: Then the Berry conjecture amounts to saying that
978: $C_A^I$ are {\it independently distributed random
979: numbers}. At this stage it is sufficient to just
980: specify the {\it two-point correlation} of this
981: distribution
982: \be \label{9}
983: \lan C_A^{*I}C_B^J\ran_{ens}= {1\over N_A}\delta_{AB}
984: \delta_{IJ}
985: \ee
986: It is easy to see that eqn (\ref{9}) is compatible
987: with unitarity. Despite the apparent basis dependence
988: of this criterion, it is actually {\em basis
989: independent}.To see this let us expand $|A\ran$ in
990: another orthonormal basis $I'$
991: \be \label{10}
992: |A\ran = \sum_I' C_A^{I'}|I'\ran
993: \ee
994: with
995: \be \label{11}
996: |I'\ran = B^{I'}_I|I\ran
997: \ee
998: Then
999: \be \label{12}
1000: C_A^{I'} = \sum_I B_I^{I'}C_A^I
1001: \ee
1002: and
1003: \br \label{13}
1004: \lan C_A^{*I'}C_B^{I'}\ran_{ens}&=&\sum_{IJ}B_I^{*I'}B_J^{J'}\lan C_A^{*I}C_B^J\ran_{ens}\nonumber\\
1005: &=&{1\over N_A}\sum_{IJ}B_I^{*I'}B_J^{J'}\delta_{AB}\delta_{IJ}\nonumber\\
1006: &=&{1\over N_A}\delta_{AB}\delta_{I'J'}
1007: \er
1008: where we have made use of the unitarity of the transformation
1009: matrix $B_I^{I'}$. It should be noted that this proof works only
1010: when the bases ${I},{I'}$ etc. are not derived from the energy basis
1011: by application of {\it fixed} unitary transformations on the
1012: energy-eigenfunction basis. Now let us consider the
1013: ensemble-average of eqn (\ref{7}):
1014: \br \label{14}
1015: \lan\sum_A |C_A^\psi|^2 O_{AA}\ran_{ens}&=&\sum_{AIJ}|C_A^\psi|^2 \lan C_A^{*I} C_A^J\ran_{ens}O_{IJ}\nonumber\\
1016: &=&{1\over N_A}\sum_{AIJ}|C_A^\psi|^2\delta_{IJ}O_{IJ}\nonumber\\
1017: &=&{1\over N_A}\sum_A|C_A^\psi|^2\sum_I O_{II}\nonumber\\
1018: &=&{1\over N_A}\sum_I O_{II}
1019: \er
1020: Let us also consider the average of $O$ at a particular instant given
1021: by eqn (\ref{4}):
1022: \br \label{15}
1023: \lan\sum_{AB}C_B^{*\psi}C_A^\psi e^{-i(E_A-E_B)t} O_{BA}\ran_{ens}
1024: &=&\sum_{ABIJ}C_B^{*\psi}C_A^\psi \lan C_A^{*I}C_B^J\ran_{ens}O_{IJ}e^{-i(E_A-E_B)t}\nonumber\\
1025: &=&{1\over N_A}\sum_{ABIJ}C_B^{*\psi}C_A^\psi \delta_{AB}\delta_{IJ}e^{-i(E_A-E_B)t}O_{IJ}\nonumber\\
1026: &=&{1\over N_A}\sum_{AI}C_A^{*\psi}C_A^\psi O_{II}\nonumber\\
1027: &=&{1\over N_A}\sum_I O_{II}
1028: \er
1029: which is the same as eqn (\ref{14}). Thus in these cases the time-average
1030: is equal to the value at any given instant of time once the ensemble
1031: average is taken
1032: and in both of 
1033: them all memory
1034: of the initial state is lost. The drawback with this picture is that the
1035: equality seems to hold at arbitrary times $t$ while one should only
1036: expect it at late times. In other words there is no way to understand
1037: "thermalization times" in this picture.
1038: 
1039: For eqn (\ref{6}) to hold
1040: it actually suffices for $T >> {1\over \delta}$ where
1041: $\delta$ is the spacing between the exact energy
1042: eigenvalues. As this spacing typically decreases
1043: exponentially with the size of the system the
1044: corresponding $T$ is exponentially large in system size.
1045: {\em This is analogous to the classical case where for
1046: very very large times the Poincare recurrence theorem
1047: \cite{arnold,chandra} guarantees the equality of time
1048: and ensemble averages.}
1049: 
1050: Clearly what's important are not such astronomically large time-scales.
1051: Instead, consider the more moderate and intermediate
1052: time-scale\\ ${1\over \Delta}~< T~<< {1\over \delta}$:
1053: then one has
1054: $$
1055: {\sin(E_m-E_n)T\over(E_m-E_n)T}\simeq 1
1056: $$
1057: and consequently
1058: \be \label{16}
1059: \bar O(\tau)=\sum C^{*\psi}_B~C^\psi_A~~e^{-i(E_A-E_B)\tau}A_{BA}
1060: \ee
1061: Again if the spectrum of the Hamiltonian is
1062: sufficiently complex and there are no degeneracies, for
1063: any reasonable value of $\tau$ there will be
1064: cancelations among all the $A\ne B$ terms and one
1065: again gets back to eqn (\ref{7}). Actually for such
1066: complex systems even the average at any particular
1067: instant(except for pathologically small values of $\tau$)
1068: becomes the same as eqn (\ref{7}) and one does not
1069: even need to do any time averaging. But eqn (\ref{7})
1070: retains memory of the initial state and at this stage
1071: it is not possible to recover the basis of statistical
1072: mechanics unless one postulates additional assumptions
1073: like Berry's conjecture etc.
1074: (we will show later ways of going beyond such
1075: restrictive assumptions; that will be the main result
1076: of this paper).
1077: 
1078:  In  \cite{sred} there was a proposal to consider a different
1079: approach to this problem by assuming that observables
1080: of interest are of the type:
1081: \be \label{17}
1082:  O_{BA}~=~O_0~\delta_{BA}~~+R_{BA}
1083: \ee
1084: with $R_{BA}$ being "small".
1085: 
1086:  Now too $\lan\psi(t)|O|\psi(t)\ran$ is of the same
1087: form as eqn(\ref{7}) except for small corrections
1088: due to the $R_{BA}$ terms. Once again if we interpret
1089: this restriction in the Schr\H{o}dinger picture
1090: one would get eqn (\ref{7}) to be true at all times
1091: which does not make much sense. A possibility is to
1092: interpret $\lan\psi(t)|O|\psi(t)\ran$ as
1093: $\lan\psi|O(t)|\psi\ran$ in the Heisenberg picture
1094: and assume eqn (\ref{17}) only for late times. Even then
1095: one has to still introduce some "ensemble average"
1096: along the lines of Berry's conjecture and there seems
1097: to be no scope for addressing the issue of
1098: "thermalization time" at all. We will see later that
1099: eqn (\ref{17}) is {\em too strong} a restriction on
1100: macroscopic observables and in fact implies that in
1101: the classical limit these observables are
1102: {\em constant} on the entire energy surface.
1103: \subsection{Earlier approaches}
1104: Much after we had finished our work( to be described
1105: in detail in secs 4, 5 and 6) we came to know of the
1106: fundamental paper on this subject due to J. von Neumann
1107: \cite{JvN} through ref \cite{GOM}. Subsequently
1108: we have traced Pauli and Fierz's \cite{pauli} sequel
1109: to von Neumann's work as well as Van Kampen's
1110: formulation \cite{kampen} of the problem. There are
1111: differences between von Neumann's and Van Kampen's
1112: approaches but they are essentially similar in spirit
1113: and both give a quantum mechanical equivalent of
1114: the classical phase space cell decomposition. While
1115: von Neumann uses fairly rigorous methods to show that
1116: the time average of the expectation values of the
1117: macroscopic observables (defined suitably by him)
1118: approaches arbitrarily closely the microcanonical
1119: average( he also shows that a suitably defined entropy,
1120: different from what is known as von Neumann entropy
1121: in the literature, also approaches the entropy of
1122: the microcanonical distribution which is the quantum
1123: version of the H-theorem). Van Kampen establishes
1124: a master equation for the probability distribution
1125: for finding the system in various phase cells given the
1126: distribution at $t=0$ and argues that it has all the
1127: properties of a Markov process whence the problem
1128: becomes identical to the one in classical statistical
1129: mechanics. While in von Neumann's treatment the issue
1130: of thermalization time is not very transparent, in
1131: Van Kampen's treatment it can be handled in principle
1132: exactly as in classical statistical mechanics.
1133: 
1134: While the
1135: mathematical formalism in all three approaches have
1136: strong overlaps,
1137: our  treatment is rather different in its
1138: interpretation of coarse graining in quantum mechanics
1139: which is somewhat kinematical in origin in both
1140: von Neumann's and Van Kampen's treatments. We seek a
1141: dynamical origin for this coarse-graining. In our approach, the
1142: extra soft quanta not only play a passive role as the
1143: degrees that are coarse grained away, they directly provide
1144: the mechanism for thermalization. Thermalization takes
1145: place due to emission and absorption of these soft quanta.
1146: Furthermore in order for this thermalization to happen
1147: it is crucial that they have the properties described
1148: in Sec 2.4 and 2.5. Thus the reader will see that
1149: the setup of Section 4,
1150: where the solution to the problem of ergodicity is
1151: given in an abstract way
1152: is, perhaps not surprisingly,
1153: mathematically identical to the setup of von Neumann. 
1154: Both of them describe a quantum mechanical analogue of Gibbsian
1155: coarse graining. Section 5 and Section 6 where the soft quanta
1156: play out their role in thermalization have, no obvious
1157: counterpart in
1158: von Neumann's discussion.
1159: 
1160: We include a short description of the works of von Neumann and van Kampen
1161: for  reasons of completeness and historical accuracy.
1162: This should also provide a perspective for the whole
1163: discussion.
1164: This section is not a prerequisite for understanding
1165: the rest of this paper. For convenience we have given
1166: at appropriate places below our notation used in Section 4
1167: corresponding to each of von Neumann's symbols.
1168: 
1169: 
1170: 
1171: 
1172: \subsubsection{Quantum Statistical Mechanics of von Neumann}
1173: The three essential ingredients of this approach are
1174: i)energy-surfaces,ii)phase cells belonging to a
1175: particular energy surface and iii)micro-states spanning
1176: each phase cell.
1177: 
1178: {\bf Energy Surfaces:}The {\em exact} Hamiltonian of the system is taken to
1179: be $H$ whose spectrum is assumed to be discrete
1180: for simplicity.If $\Delta E$ is the macroscopic
1181: resolution with which energy measurements can be made,
1182: the energy levels are divided into groups of width
1183: $\Delta E$ each and the groups are labeled by
1184: $a=1,2,...$. This means that only energy levels
1185: belonging to different groups are macroscopically
1186: distinct.The energy eigenvalues are labeled
1187: $W_{\rho,a}$ with $\rho = 1,2,...,S_a$.
1188: (In Section 4 we have used $A$ to label the exact energy
1189: eigenstates. $S_a$ corresponds to $N_A$ of Section 4, 
1190: where we
1191: have restricted ourselves to one energy surface.)
1192: The eigenfunctions are labeled $\phi_{\rho,a}$. Thus
1193: $S_a$ is the number of {\em microscopic} states
1194: spanning the energy surface $a$. The projection operator
1195: corresponding to $\phi_{\rho,a}$ is denoted by
1196: ${\bf P}_{\phi_{\rho,a}}$.
1197: 
1198: {\bf Macroscopic Observables:}
1199: von Neumann considers all macroscopic observations
1200: to be simultaneous and therefore {\em all} macroscopic
1201: observables to be {\em mutually commuting}.
1202: (The word ``macroscopic'' is used here in the sense
1203: of Quantum Measurement Theory. It does {\em not}
1204: mean macroscopic in the sense of statistical mechanics
1205: (i.e. pressure, density, etc ).
1206: Indeed they correspond to {\em micro-state} variables
1207: such as the position coordinates of molecules of a gas
1208: except that {\em macroscopic measurements} can not 
1209: resolve their values within a phase cell.
1210: These observables are thus a subset of  the $O$  of Section 4).
1211: Their
1212: {\em simultaneous} eigenfunctions are denoted by
1213: $\omega_{\lambda,p},\lambda=1,2,..,s_p$ where distinct
1214: values of $p$ denote distinct values of the observables
1215: but all states with the same $p$ but different
1216: $\lambda$ have the same values for all observables.
1217: (In Section 4 we have used `$a$` for $\lambda$, and
1218: `$i$` for $p$. $s_p$ is $N_a$ of Section 4. The simultaneous eigenstates
1219: there are denoted by $| i,a \rangle$.)
1220: 
1221: The projection operator for the state
1222: $\omega_{\lambda,p}$ is denoted by
1223: ${\bf P}_{\phi_{\lambda,p}}$ and the density matrix
1224: signifying an equal mixture of $\omega_{\lambda,p}$
1225: for the same $p$ but for all possible values of
1226: $\lambda$ is denoted by
1227: \be \label{18}
1228: {1\over s_p} {\bf E}_p \equiv \sum^{s_p}_{\lambda}{\bf P}_{\omega_{\lambda,p}}
1229: \ee
1230: Eqn (\ref{18}) can be taken as the density matrix for
1231: a "phase space cell".
1232: In von Neumann's paper the explicit construction of
1233: these mutually commuting macroscopic observables is
1234: not given and the discussion not very transparent.
1235: Van Kampen treats this in a more transparent manner.
1236: Consider any observable $O$ and express it in the
1237: exact energy-eigenfunction basis
1238: \be \label{19}
1239: O=\sum_{\rho,a,\sigma,b} O_{\rho,a;\sigma,b}|\phi_{\rho,a}\ran\lan\phi_{\sigma,b}|
1240: \ee
1241: Van Kampen argues that the matrix elements of the
1242: time-dependent operator $O(t)$
1243: \be \label{20}
1244: O(t)_{\rho,a,\sigma,b}=O_{\rho,a,\sigma,b}e^{-i(W_{\rho,a}-W_{\sigma,b})}
1245: \ee
1246: are associated with {\em very rapid fluctuations}
1247: whenever $a\ne b$.Thus it makes sense to replace the
1248: microscopic operator $O$ by
1249: \be \label{21}
1250: O'=\sum_{\rho,a,\sigma,a} O_{\rho,a;\sigma,a}|\phi_{\rho,a}\ran\lan\phi_{\sigma,a}|
1251: \ee
1252: Equivalently
1253: \be \label{22}
1254: O'_{\rho,a;\sigma,b}=O_{\rho,a;\sigma,a}~\delta_{ab}
1255: \ee
1256: Van Kampen also introduces the coarse-grained energy
1257: operator to be
1258: \be \label{23}
1259: H' = \sum_a W_a \sum_\rho^{S_a}{\bf P}_{\phi_{\rho,a}}
1260: \ee
1261: Note that all the energy eigenvalues belonging to
1262: an energy-surface have been replaced by a single
1263: value $W_a$. Now it is easy to see that $H',O'$
1264: commute so they can be
1265: {\em simultaneously diagonalised}:
1266: \br \label{24}
1267: H'|\omega_{\mu,a}\ran &=& W_a|\omega_{\mu,a}\ran\nonumber\\
1268: O'|\omega_{\mu,a}\ran &=& A_{\mu,a}|\omega_{\mu,a}\ran
1269: \er
1270: He now introduces the coarse-grained observables by
1271: grouping the eigenvalues $A_{\mu,a}$ into phase cells
1272: in such a way that
1273: only those belonging to different groups(cells) can be
1274: distinguished macroscopically i.e through macroscopic
1275: measurements.Let these groupings be labeled by
1276: $\nu=1,2,...N_a$ ($N_a$ of von Neumann is $N_i$ of Section 4). Then
1277: \be \label{25}
1278: {\tilde O} = \sum_a\sum_\nu A_{\nu,a}\sum_{\lambda}^
1279: {s_{\nu,a}}{\bf P}_{\omega_{\lambda,\nu,a}}
1280: \ee
1281: Now it is clear that all such macroscopic observables
1282: commute with each other with $\omega_{\lambda,\nu,a}$
1283: as their simultaneous eigenfunctions. Furthermore,in
1284: this approach, the phase cells, labeled by $\nu,a$
1285: are neatly partitioned into disjoint sets by the
1286: energy surfaces.
1287: 
1288: It is instructive to compare this construction with
1289: von Neumann's treatment( and later with ours). von
1290: Neumann argues that any operator with a boolean
1291: spectrum( eigenvalues $0$ or $1$) is a macroscopic
1292: observable and every macroscopic observable has
1293: a spectral decomposition of the type
1294: \be \label{26}
1295: A=\sum_p c_p {\bf E}_p
1296: \ee
1297: He then considers the function $f_a(x)$ such that $f_a=1$
1298:  if $x$ belongs to the energy eigenvalues of the
1299: energy surface labeled by $a$ and $0$ otherwise. Then
1300: the operator $f_a(H)$ has eigenvalues $0,1$ and is
1301: thus a macroscopic observable. By eqn(\ref{26}) it
1302: must admit the decomposition
1303: \be \label{27}
1304: f_a(H)=\sum_p f_a^p {\bf E}_p
1305: \ee
1306: On the other hand
1307: \be \label{28}
1308: f_a(H)=\sum_{\rho=1}^{S_a}{\bf P}_{\phi_{\rho,a}}
1309: \ee
1310: As both $f_a(H)$ and ${\bf E}_p$ for every p are equal to
1311: their own squares and as the product of two distinct
1312: ${\bf E}_p$'s is zero, it follows that ${f_a^p}^2 = {f_a^p}$
1313: for every $p$ i.e $f_a^p$ is either $1$ or $0$.
1314: Relabeling all those $p$'s for which$f_a^p=1$ by
1315: $\nu,a$ we have
1316: \be \label{29}
1317: \sum_\rho^{S_a} {\bf P}_{\phi_{\rho,a}} = \sum_\nu^{N_a} {\bf E}_{\nu,a}
1318: \ee
1319: i.e we again have a unique partitioning of all the
1320: phase cells among the energy surfaces. The equivalent 
1321: of this in our approach is given by the construction in sec. 4.1.
1322: 
1323: {\bf Microcanonical ensemble:}\\
1324: We briefly sketch JvN's proof showing the asymptotic
1325: equivalence of the time averaged expectation value and
1326: the microcanonical average.
1327: We denote the operator occurring in eqn (\ref{29}) by
1328: $\Delta_a$. It is then clear that
1329: ${1\over S_a}\Delta_a$ is the density matrix for
1330: an equal mixture of energy eigenstates belonging
1331: to the energy surface $a$ and as such represents
1332: the density matrix for the microcanonical ensemble associated
1333: with that energy surface. If the initial state $\psi$ is such that
1334: $(\Delta_a\psi,\psi)$ is nonzero for several values of $a$, one
1335: should consider a mixture of microcanonical ensembles for each $a$
1336: with weight factor $(\Delta_a\psi,\psi)$(this is the probability
1337: of finding the system in the energy surface $a$). Thus the microcanonical
1338: density matrix in the general case is
1339: \be \label{30}
1340: U_\psi = \sum_{a=1}^\infty {(\Delta_a\psi,\psi)\over S_a}\Delta_a
1341: \ee
1342: In our treatment later on we shall consider only one energy surface.
1343: 
1344: {\bf Quantum Ergodicity}
1345: 
1346: Next JvN considers the system to be initially in a state
1347: \be \label{3.2.31}
1348: |\psi\ran = \sum_{a=1}^\infty\sum_{\rho=1}^{S_a} r_{\rho,a}e^{i\alpha_{\rho,a}}
1349: |\phi_{\rho,a}\ran ~~~~(r_{\rho,a}\geq 0,0\leq\alpha_{\rho,a}< 2\pi)
1350: \ee
1351: then
1352: \be \label{3.2.32}
1353: |\psi(t)\ran = \sum_{a=1}^\infty\sum_{\rho=1}^{S_a}r_{\rho,a}e^{-i({2\pi\over h}
1354: W_{\rho,a}t-\alpha_{\rho,a})}
1355: \ee
1356: Introducing the abbreviations
1357: \be \label{3.2.33}
1358: x_{\nu,a}=({\bf E}_{\nu,a}\psi(t),\psi(t))~~~~u_a=(\Delta_a\psi(t),\psi(t))=(\Delta_a\psi,\psi)
1359: \ee
1360: one considers a macroscopic observable $A$
1361: \be \label{3.2.34}
1362: A=\sum_{a=1}^\infty\sum_{\nu=1}^{N_a}~\eta_{\nu,a}{\bf E}_{\nu,a}
1363: \ee
1364: The expectation value of $A$ in the state $\psi(t)$ is then
1365: \be \label{35}
1366: E_A(\psi(t))=(A\psi(t),\psi(t))= \sum_{a=1}^\infty\sum_{\nu=1}^{N_a
1367: } \eta_{\nu,a}({\bf E}_{\nu,a}\psi(t),\psi(t))
1368: =\sum_{a=1}^\infty\sum_{\nu=N_a}\eta_{\nu,a}x_{\nu,a}
1369: \ee
1370: while the microcanonical average is
1371: \br \label{36}
1372: E_A(U_\psi)=Tr AU_\psi &=&Tr(\sum_{a=1}^\infty\sum_{\nu=1}^{N_a}\eta_{\nu,a}{\bf E}_{\nu,a}
1373: \sum_{a'=1}^\infty\sum_{\nu'=1}^{N_{a'}}{u_{a'}\over S_{a'}}{\bf E}_{\nu',a'})\nonumber\\
1374: &=&\sum_{a=1}^\infty\sum_{\nu=1}^{N_a}\eta_{\nu,a}{u_a s_{\nu,a}\over
1375:  S_a}
1376: \er
1377: On using the eqn (\ref{36}) and Schwarz's inequality JvN finally obtains
1378: \be \label{37}
1379: (E_A(\psi(t))-E_A(U_\psi))^2\leq {\bar \eta}^2\sum_{a=1}^\infty\sum_{\nu=1}^{N_a}
1380: {S_a\over s_{\nu,a}u_a}[x_{\nu,a}-{s_{\nu,a}u_a\over S_a}]^2
1381: \ee
1382: where
1383: \be \label{38}
1384: {\bar \eta}^2 = \sum_{a=1}^\infty\sum_{\nu=1}^{N_a}{s_{\nu,a}u_a\over S_a}\eta_{\nu,a}
1385: \ee
1386: JvN finally establishes that the time average of the lhs of eqn (\ref{37})
1387: is bounded by ${\bar \eta}^2 max_a\{{2N_a\over S_a}\}$.
1388: \subsubsection{Van Kampen's approach}
1389: Van Kampen expands the initial state $\psi$ not in terms
1390: of the exact eigenfunctions of the Hamiltonian but
1391: in terms of the simultaneous eigenstates
1392: $\omega_{\lambda,\nu,a}$ of the coarse grained
1393: observables( we are retaining the notation of JvN):
1394: \be \label{39}
1395: |\psi\ran = \sum_{\lambda,\nu,a} b_{\lambda,\nu,a}|\omega_{\lambda,\nu,a} \ran
1396: \ee
1397: The probability of finding the system in the phase cell
1398: $\nu,a$ is given by
1399: \be \label{40}
1400: P_{\nu,a}=\sum_{\lambda=1}^{s_{\nu,a}}|b_{\lambda,\nu,a}|^2
1401: \ee
1402: The time dependence of the state is given by
1403: \be \label{41}
1404: |\psi(t)\ran = \sum_{\lambda,\nu,a}b_{\lambda,\nu,a}(t)|\omega_{\lambda,\nu,a}\ran
1405: \ee
1406: where
1407: \be \label{42}
1408: b_{\lambda,\nu,a}(t)=\sum_{\lambda',\nu',a'}\lan \omega_{\lambda,\nu,a}|U(t)|\omega_{\lambda',\nu',a'}\ran
1409: \ee
1410: and $U(t)$ is the time-evolution operator. The
1411: probability of finding the system in the phase cell $\nu,a$ at time $t$ is given by
1412: \be \label{43}
1413: P_{\nu,a}(t) = \sum_{\stackrel{\lambda',\nu',a'}{\lambda'',\nu'',a''}}
1414: \sum_{\lambda}\lan\omega_{\lambda,\nu,a}|U(t)|\omega_{\lambda',\nu',a'}\ran
1415: \lan \omega_{\lambda,\nu,a}|U(t)|\omega_{\lambda'',\nu'',a''}\ran^*
1416: b_{\lambda',\nu',a'}(0)b^*_{\lambda'',\nu'',a''}(0)
1417: \ee
1418: Invoking the argument that the summation consists of many wildly fluctuating
1419: terms and all the non-negative terms cancel i.e only surviving terms are
1420: ${\nu',a'}={\nu'',a''}$ and $\lambda'=\lambda''$.This reduces eqn(\ref{41})
1421: to
1422: \be \label{44}
1423: P_{\nu,a}(t)=\sum_{\nu',a'}\sum_{\lambda'}\sum_{\lambda}
1424: |\lan\omega_{\lambda,\nu,a}|U(t)|\omega_{\lambda',\nu',a'}\ran|^2 |b_{\lambda',\nu',a'}(0)|^2
1425: \ee
1426: Van Kampen further argues that in a sum of type
1427: $$
1428: \sum_i\alpha_i\beta_i
1429: $$
1430: where both $\alpha_i,\beta_i$ are {\it rapidly fluctuating} but {\it positive}
1431: it is a good approximation to evaluate the sum as
1432: $$
1433: {1\over G}\sum_i\alpha_i\sum_i\beta_i
1434: $$
1435: With this approximation eqn (\ref{44}) becomes
1436: \be \label{45}
1437: P_{\nu,a}(t) = \sum_{\nu',a'}{1\over s_{\nu,a}}
1438: \sum_{\lambda,\lambda'}
1439: |\lan\omega_{\lambda,\nu,a}|U(t)|\omega_{\lambda',\nu',a'}\ran|^2
1440: P_{\nu',a'}(0)
1441: \ee
1442: With the definition
1443: \be \label{46}
1444: T_t(\nu,a|\nu',a') {\stackrel {def}{=}}
1445: {1\over s_{\nu,a}}\sum_{\lambda,\lambda'}
1446: |\lan\omega_{\lambda,\nu,a}|U(t)|\omega_{\lambda',\nu',a'}\ran|^2
1447: \ee
1448: eqn (\ref{45}) can be recast as
1449: \be \label{47}
1450: P_{\nu,a}(t) = \sum_{\nu',a'}T_t(\nu,a|\nu',a') P_{\nu',a'}(0)
1451: \ee
1452: As there is nothing special about the instant $t=0$ it follows that
1453: \be \label{48}
1454: P_{\nu,a}(t_1+t_2) = \sum_{\nu',a'}T_{t_2}(\nu,a|\nu',a') P_{\nu',a'}(t_1)
1455: \ee
1456: Equivalently
1457: \be \label{49}
1458: T_{t_1+t_2}(\nu,a|\nu',a')=\sum_{\nu'',a''} T_{t_2}(\nu,a|\nu'',a'')T_{t_1}
1459: (\nu'',a''|\nu',a')
1460: \ee
1461: Along with the obvious properties of $T_t(\nu,a|\nu',a')$
1462: \be \label{50}
1463: T_0(\nu,a|\nu',a')=\delta_{\nu,a;\nu',a'};~~~T_t(\nu,a|\nu',a')\geq 0;~~~
1464: \sum_{\nu,a}T_t(\nu,a|\nu',a')=1
1465: \ee
1466: which follow
1467: from the orthogonality of $|\omega_{\lambda,\nu,a}\ran,|\omega_{\mu,\nu',a'}\ran$
1468: for distinct $\{\nu,a\},\{\nu',a'\}$ and the {\em unitarity} of $U(t)$, one
1469: concludes that $T_t$ is a {\em Markov Process}. One knows from the
1470: Frobenius-Perron theorem that there is always an equilibrium distribution
1471: $P^{eq}_{\nu,a}$. From the spectrum of the operator $T(\nu,a|\nu',a')$ one
1472: can also deduce the {\em thermalization times}.
1473: 
1474: Before concluding this section it is instructive to derive the so called
1475: Master equation. For this one solves for $T_{\Delta t}$ as
1476: \be \label{51}
1477: T_{\Delta t}(\nu,a|\nu',a')=\delta_{\nu,a;\nu',a'}\{1-\Delta t\sum_{\nu'',a''}
1478: W_{\nu'',a'';\nu,a}\}+\Delta t W_{\nu,a;\nu',a'}
1479: \ee
1480: Then one gets the differential form of the Chapman-Kolmogorov equation
1481: \be \label{52}
1482: {\partial\over \partial t}T_t(\nu,a|\nu',a')=
1483: \sum_{\nu'',a''} W_{\nu,a;\nu'',a''}T_t(\nu'',a''|\nu',a')-W_{\nu'',a'';\nu,a}
1484: T_t(\nu,a|\nu',a")
1485: \ee
1486: This can be recast in terms of $P_{\nu,a}(t)$ as
1487: \be \label{53}
1488: {d\over dt}P_{\nu,a}(t)=\sum_{\nu',a'}\{W_{\nu,a;\nu',a'}P_{\nu',a'}
1489: -W_{\nu',a';\nu,a}P_{\nu,a}(t)\}
1490: \ee
1491: which is the {\em master equation}.
1492: \section {Quantum coarse graining and emergence of the microcanonical
1493: ensemble.}
1494: \setcounter{equation}{0}
1495: 
1496: As we have stated already the major conceptual problem in Quantum statistical
1497:  mechanics
1498: is that of understanding how a quantum state initially described by a pure 
1499: density
1500: matrix can ever resemble the state of thermodynamic equilibrium which is
1501: clearly
1502: described by a mixed density matrix.
1503: \subsection{Coarse graining}
1504: As was clearly stressed by Gibbs \cite{gibbs} long ago, coarse graining is 
1505: essential to understanding
1506: even the most basic features of classical statistical mechanics. We now introduce
1507: our proposal
1508: for a {\em quantum coarse graining} and show subsequently that the microcanonical
1509: distribution then follows for a large class of systems.
1510: 
1511: We consider two types of mutually commuting variables $I,{\cal A}$ and their
1512: simultaneous eigenstates
1513: $|i,a\ran $. Later on we will define $|i,a\ran $ only in the {\em subspace spanning the
1514: states
1515: relevant for the microcanonical ensemble}. The $i$ will eventually correspond to
1516: the labels of the usual micro-states entering
1517: the description of the statistical system and the precise nature of $a$ are not
1518: specified
1519: at present. Alternately one can think of $|i,a\ran $ as the labeling of micro-states on
1520: the fine scale and
1521: $|i\ran $ as the labeling of the micro-states at the coarse grained level.
1522: 
1523: It is assumed that to a very good approximation the 
1524: ${\cal A}$ variables couple
1525: {\em weakly}
1526: to the $I$-variables in the total Hamiltonian i.e
1527: \be \label{4.1}
1528: H_{tot}=H^{(i)}+H^{(a)}+\lambda~~H^{(i,a)}
1529: \ee
1530: with $\lambda$ very small. Let $|A^*\ran $ be the eigenstates of $H^{(i)}$ {\bf that
1531: lie in
1532: the range $E_0-\Delta E<E_0+\Delta$}. We then form the
1533: matrix $I^*_{AB} = \lan B^*|I|A^*\ran $ which is $N_i\times N_i$. Now let $|i\ran $ be the
1534: eigenstates of
1535: $I^*$ and let $|i,a\ran= |i\ran \otimes|a\ran $ where $|a\ran $ are the eigenstates of $\cal A$
1536: and let
1537: us also assume that there are $N_a$ such states. (A special choice
1538: could be $I=H^{(i)}$, in which case $|i\rangle = |A^* \rangle $.)
1539: 
1540: The important point is that the $N_i\times N_a$ states $|i,a\ran $ continue to form a
1541: basis in terms of
1542: which all the exact energy eigenstates in the microcanonical band can still be
1543: expanded. It is
1544: of course important to state that even a small perturbation characterized by
1545: $\lambda$ can completely
1546: alter the nature of the eigenstates of the Hamiltonian $H$ because of the exponential
1547: crowding of the "unperturbed" states but the {\em number} of eigenstates remains
1548: unchanged.
1549: The parameters of the microcanonical band $E_0,\Delta$ are also most likely
1550: affected by
1551: the perturbation, but as we shall never need their precise values it does not matter
1552: for the ensuing discussion.
1553: 
1554: We shall only consider those observables that are {\em insensitive} to the $a$
1555: labels i.e $\lan i,a|O|j,b\ran
1556: =O_{ij}\delta_{ab}$. This is in keeping with the spirit of
1557: coarse graining described in Section 2.2.
1558: 
1559: \subsection{Some preliminaries}
1560: Consider some arbitrary state in the subspace we have considered
1561: \be \label{4.2}
1562: |\psi\ran =\sum C^{\psi}_A |A\ran
1563: \ee
1564: where now $|A\ran $ are the {\em exact} energy eigenstates of the system.
1565: Further, let
1566: \be \label{4.3}
1567: |A\ran  = \sum_{i,a}~~C_{A}^{*ia}~~|i,a\ran
1568: \ee
1569: with the inverse expansion
1570: \be \label{4.4}
1571: |i,a\ran  = \sum_A~~C^{ia}_A~~|A\ran
1572: \ee
1573: 
1574: The coefficients $C_A^{ia}$ satisfy the following unitarity conditions:
1575: \be \label{4.5}
1576: \sum_A C_A^{*ia}~C_A^{i'a'}=\delta_{ii'}\delta_{aa'}~~~~~~\sum_{ia} C_A^{ia}C_{A'}^{*ia}=\delta_{AA'}
1577: \ee
1578: 
1579: 
1580: In terms of these definitions we can rewrite
1581: \be \label{4.6}
1582: \lan \psi|O|\psi\ran  = \sum_{A,B}~~C^{*\psi}_B~~C^{\psi}_A~~\lan B|O|A\ran
1583: \ee
1584: as
1585: \be \label{4.7}
1586: O_{\psi\psi}~~=~~\sum_{a,ij}~\sum_{A,B}~C^{*\psi}_B~~C^{\psi}_A~~
1587: C_{A}^{*ia} ~ C_{B}^{ja} O_{ij}
1588: \ee
1589: Now introduce
1590: \be \label{4.8}
1591: \sum_a C_{A}^{ja} C_{B}^{*ka}=~~\delta_{jk}~~\delta_{AB}~~P^j_A+R^{jk}_{AB}
1592: \ee
1593: This is to be understood as a {\em diagonal $+$ off-diagonal} split in
1594:  the sense
1595: that by
1596: definition
1597: \be \label{4.9}
1598: P^j_A~~=~~\sum_b C_{A}^{jb}~C_{A}^{*jb}
1599: \ee
1600: Consequently
1601: \be \label{4.10}
1602: R^{kk}_{AA}=0
1603: \ee
1604: (No sum on indices.)
1605: We need to prove a few important properties of $P$ and $R$. Let $A\ne B$;
1606:  putting $j=k$ and
1607: summing over j one gets
1608: \be \label{4.11}
1609: \sum_k R^{kk}_{AB} = 0
1610: \ee
1611: Therefore
1612: \be \label{4.12}
1613: \sum_k R^{kk}_{AB}=0
1614: \ee
1615: always. Further,
1616: \be \label{4.13}
1617: \sum_k~~P^k_A~~=1  ~and ~~ also~ \sum_A P^k_A=N_a
1618: \ee
1619: as follow from the unitarity eqn (\ref{4.5}).  Likewise by putting $A=B$ and summing
1620: over $A$ one gets
1621: \be \label{4.14}
1622: \sum_A~~R^{kj}_{AA}=0
1623: \ee
1624: 
1625: 
1626: \subsection{Emergence of the microcanonical distribution}
1627: Let us consider a pure quantum state which at $t=0$ is given by
1628: \be \label{4.15}
1629: |t=0\ran ~=~|\psi\ran ~=~\sum_A~~C^{\psi}_A~~|A\ran
1630: \ee
1631: Now it is straightforward to show that
1632: \be \label{4.16}
1633: \lan \psi(t)|O|\psi(t)\ran ~~=~~\sum_{A,j}~~C^{*\psi}_A~~C^{\psi}_A~~P^j_A~~O_{jj}+
1634: \sum_{ABjk}~~C^{*\psi}_B~C^{\psi}_A~E^{-i(E_A-E_B)t}~~R^{kj}_{AB}~O_{jk}
1635: \ee
1636: Separating out the $A=B$ contribution in the second term one finds
1637: \br  \label{4.17}
1638: \lan \psi(t)|O|\psi(t)\ran ~~=~~& &\sum_{A,j}~~C^{*\psi}_A~~C^{\psi}_A~~P^j_A~~O_{jj}
1639:  +\sum_A~~C^{*\psi}_A~C^{\psi}_A\sum_{jk} R^{kj}_{AA}~O_{jk}\nonumber\\
1640: &+& \sum_{A\ne B}~~\sum_{jk}C^{*\psi}_B~
1641: C^{\psi}_A~E^{-i(E_A-E_B)t}~~R^{kj}_{AB}~O_{jk}
1642: \er
1643: 
1644: We can further exploit the redundancy introduced by $a$ states by
1645: defining an equivalence class of states as follows:
1646: define the unitary operator $U^{\alpha}$ by
1647: \be \label{4.18}
1648: U^{\alpha}|a\ran = \sum _b U^{\alpha}_{ab}|b\ran
1649: \ee
1650: which is some unitary transformation on $|a\ran$ space. Clearly
1651: it does not affect the expectation values of our observables of interest, $O$,
1652: since $O$ is insensitive to $a$. Thus if
1653: \be \label{4.19}
1654: U^{\alpha}|\psi(t) \ran = |\psi ^{\alpha}(t)\ran
1655: \ee
1656: we have:
1657: \be \label{4.20}
1658: \lan \psi ^{\alpha}(t)|O|\psi ^{\alpha}(t) \ran = \lan \psi(t) |O|\psi(t) \ran
1659: \ee
1660: Consequently
1661: \be \label{4.21}
1662: \lan \psi (t) |O|\psi (t) \ran = {1\over N_{\alpha}}\sum _{\alpha}
1663: \lan \psi ^{\alpha} (t) |O|\psi ^{\alpha}(t) \ran
1664: \ee
1665: Here $N_{\alpha}$ is the number of such $U$ matrices. This is up to us and we
1666: can choose these matrices to satisfy
1667: \be   \label{4.22}
1668: \sum _{\alpha}U^{\alpha}_{ac} U^{* \alpha}_{bd}= \mu \delta _{ab}
1669: \delta _{cd}
1670: \ee
1671: If we multiply by $\delta _{ab}$ and sum over $a,b$, we get, using the
1672: unitarity of the matrix,
1673: \[
1674: \sum _\alpha \delta _{cd} = \mu N_a \delta _{cd}
1675: \]
1676: This gives $N_\alpha = \mu N_a$.  The conditions (\ref{4.22}) are
1677: $N_a^4$ conditions on $N_\alpha N_a^2$ variables. Thus we need $N_\alpha = N_a^2$ at least. So we take $N_\alpha = N_a^2$ which gives $\mu = N_a$.
1678: 
1679: It should be noted that eqn (\ref{4.21}) is valid {\em only at a particular
1680: instant $t$} as it can be valid for {\em all $t$} only if $U$ commutes
1681: with the microscopic Hamiltonian $H$. We can think of $|\psi^{\alpha}(t)\ran$
1682: as time evolved (after $t$) from some state $|\psi^{\alpha}\ran$. In other
1683: words
1684: \be \label{4.23}
1685: |\psi^{\alpha}(t)\ran=e^{-iHt}|\psi^{\alpha}\ran
1686: \ee
1687: Again eqn (\ref{4.22}) is only valid for time $t$. Actually $|\psi^{\alpha}\ran$
1688: explicitly depends on $t$ unless $[U,H]=0$.
1689: Let
1690: \be \label{4.24}
1691: |\psi^{\alpha}\ran=\sum_A C_A^{\psi^{\alpha}}|A\ran
1692: \ee
1693: Then eqn(\ref{4.21}) implies
1694: \be \label{4.25}
1695: \sum_A C_A^{\psi^{\alpha}} e^{-iE_At}|A\ran=U_{\alpha}\sum_B C_B^{\psi}e^{-iE_Bt}|B\ran
1696: \ee
1697: Fromthis it follows that
1698: \be \label{4.26}
1699: C_A^{\psi^{\alpha}}=\sum_B e^{i(E_A-E_B)t}\lan A|U_{\alpha}|B\ran C_B^{\psi}
1700: \ee
1701: Let us evaluate $\lan A|U_{\alpha}|B\ran$ now:
1702: \br \label{4.27}
1703: \lan A|U_{\alpha}|B\ran &=& \sum_{ijab} C_B^{*jb}C_A^{ia}\lan ia|U_{\alpha}| jb \ran \nonumber\\
1704: &=&\sum_{iab} C_B^{*ib} C_A^{ia} U_{\alpha}^{ba}
1705: \er
1706: Thus
1707: \be \label{4.28}
1708: C_A^{\psi^{\alpha}}=\sum_{Biaa'} e^{i(E_A-E_B)t} C_B^{*ia'} C_A^{ia} U_{\alpha}^{a'a} C_B^{\psi}
1709: \ee
1710: On introducing
1711: \be \label{4.29}
1712: |\psi(t)\ran = \sum_{ia} D_{ia}^{\psi}(t)|ia\ran
1713: \ee
1714: it is easy to see that
1715: \be \label{4.30}
1716: D_{ia}^{\psi}(t)=\sum_B C_B^{*ia} C_B^{\psi} e^{-iE_Bt}
1717: \ee
1718: hence
1719: \be \label{4.31}
1720: C_A^{\psi^{\alpha}}=\sum_{iaa'} e^{iE_At} C_A^{ia} D_{ia'}^{\psi}(t) U_{\alpha}^{a'a}
1721: \ee
1722: 
1723: We now use all this in equation (\ref{4.21}).
1724: \be \label{4.32}
1725: {1\over N_{\alpha}}\sum _{\alpha}
1726: \lan \psi ^{\alpha} (t) |O|\psi ^{\alpha}(t) \ran
1727: = {1\over N_\alpha}\sum_{\alpha ABkle} C^{*\psi ^\alpha}_B
1728: C^{\psi ^\alpha}_A C^{*ke}_AC^{le}_B O_{kl}e^{-i(E_A-E_B)t}
1729: \ee
1730: On using eqn (\ref{4.31})
1731: \br \label{4.33}
1732: e^{-i(E_A-E_B)t}\sum_{\alpha} C_B^{*\psi^\alpha}C_A^{\psi^\alpha}
1733: &=&\sum_{ijacbd} D^\psi _{ic}(t)D^{*\psi}_{jd}(t)\sum _\alpha
1734: U_\alpha ^{ca}U_{*\alpha}^{db} C_A^{ia}C^{*jb}_B\nonumber\\
1735: &=&\mu\sum_{ijac} D^\psi _{ic}(t)D^{*\psi}_{jc}(t) C_A^{ia}C^{*ja}_B\nonumber\\
1736: &=&\mu\sum_{ijc} D^\psi _{ic}D^{*\psi}_{jc}[P_A^i\delta_{AB}\delta_{ij}+R_{AB}^{ij}]  
1737: \er
1738: where we have made use of eqn(\ref{4.30}) and eqn(\ref{4.8}).
1739: Plug this into (\ref{4.32}) to get
1740: \be \label{4.34}
1741: {\mu\over N_\alpha} \sum_{ijklcAB}D^\psi _{ic}(t)D^{*\psi}_{jc}(t)~[P_A^i \delta _{AB}\delta _{ij}+ R_{AB}^{ij}]
1742: [P_A^k\delta _{AB}\delta _{kl}+ R_{BA}^{lk}]O_{kl}
1743: \ee
1744: %Once again all indices are summed over.
1745: 
1746: We expand to get
1747: \[
1748: {1\over N_a}\sum_A [ D^\psi _{ic}D^{*\psi}_{ic}P_A^iP_A^kO_{kk}+
1749: D^\psi _{ic}D^{*\psi}_{jc}R_{AA}^{ij}P_A^kO_{kk} + D^\psi _{ic}D^{*\psi}_{ic}
1750: P_A^i R_{AA}^{kl}O_{kl}]
1751: \]
1752: \be \label{4.35}
1753: +{1\over N_a}\sum_{A\ne B}D^\psi _{ic}D^{*\psi}_{jc}R_{AB}^{ij}R_{AB}^{kl}O_{kl}
1754: \ee
1755: Summation over all other indices as well as the time dependence of
1756: $D^\psi_{ia}$'s is understood.
1757: 
1758: If $P^j_A$  are such that there is a {\em very weak}
1759: dependence on either $j$ or $A$, we can draw
1760: the following additional conclusions:
1761: \be \label{4.37}
1762: P^j_A~=~{1\over N_i}
1763: \ee
1764: This follows from eqn(\ref{4.13})
1765: 
1766: The following is one of the ways of realising eqn (\ref{4.37}): $C_A^{ia}C_A^{*ia}$
1767: is the overlap between $|i,a\ran$ and $|A\ran$ states. If a state
1768: $|i,a\ran$ has equal amounts of $|A\ran$ states then  $C_A^{ia}C_A^{*ia}$
1769: will be more or less independent of $A$. This is equivalent to saying that
1770:  the perturbation $H^{i,a}$ ``thoroughly mixes'' up the eigenstates.
1771: Of course in
1772: general if we plot  $C_A^{ia}C_A^{*ia}$ as a function of $A$,
1773: while we expect it to be independent of $A$ on the average over a range,
1774:  we also
1775: expect the plot to be full of spikes and  dips.
1776: The sum over $a$ smoothes out these fluctuations and should
1777: produce a smooth constant function.
1778: This is similar in spirit to the Berry conjecture 
1779: though much weaker than it. 
1780: 
1781: On noting that $\sum_A R_{AA}^{ij}=0$, the final result for
1782: $\lan\psi(t)|O|\psi(t)\ran$ is
1783: 
1784: \be \label{4.38}
1785: \lan \psi(t)|O|\psi(t)\ran ~~~~{1\over N_i}~~\sum_j~~O_{jj}
1786: +{1\over N_a}\sum _{i,j,k,l,c,A\ne B}D^\psi _{ic}(t)D^{*\psi}_{jc}(t)R_{AB}^{ij}R_{AB}^{kl}O_{kl}
1787: \ee
1788: 
1789: We have used the fact that the sum over $A$ gives a factor of $N_aN_i$
1790: in the first term.  Note that the time dependent part has an explicit
1791: $1\over N_a$ in front.
1792: 
1793: It is imperative to show that the second term is
1794: negligibly small compared to the first if we are to
1795: demonstrate the emergence of the microcanonical
1796: ensemble.
1797: The spectrum of the Hamiltonian for a typical macro-system will be so
1798: complex that the second term
1799: will go to $0$ for large times. Then we have our desired result.
1800: 
1801: It is also important to know how fast the second term
1802: vanishes and for this purpose some estimation of the
1803: magnitude of the second term is important.
1804: 
1805: When the expansion coefficients $C_A^{ia}$ are
1806: identically distributed independent random variables
1807: such an estimation can be carried out quite accurately.
1808: In accordance with general ideas
1809: expressed in \cite{Berry,sred}, such a
1810: circumstance could arise when the quantum system is
1811: {\em classically chaotic}.
1812: The nature of the eigenstates then
1813: is such that at high enough energies, all eigenstates 
1814: look more or less the same. But it should be emphasised that our 
1815: approach does not really require any strong
1816: statement like this but only the weaker statement of eqn (\ref{4.38}).
1817: 
1818: The assumption about wildly fluctuating phases
1819: allows us to estimate the magnitude of $R_{AB}^{ij}$. The magnitude
1820: of any given $C_A^{ia}$ can be estimated to be
1821: $1\over \sqrt {N_iN_a}$.
1822: Thus $R_{AB}^{ij}$ is a sum of $N_a$ terms each of magnitude
1823: $1\over N_i N_a$ and random phase. Thus we expect it to be $O({{\sqrt N_a }
1824: \over N_a N_i})\approx {1\over N_i \sqrt N_a }$. Also $P_A^i \approx {1\over
1825: N_i}$.  The off diagonal part  thus has an extra factor of
1826: $1\over \sqrt N_a$ relative to the diagonal part. 
1827: 
1828: 
1829: We can similarly estimate the off-diagonal part in (\ref{4.38}).
1830: We have
1831: \[
1832: \lan \psi(t)| O | \psi (t)\ran =
1833: \sum _{j,k.A,B,b} C_A^{*\psi}C_B^{\psi}C_A^{jb}C_B^{*kb}e^{-i(E_B-E_A)t}O_{jk}
1834: \]
1835: 
1836: The off diagonal part is when $A \ne B$. There are $\approx N_A^2$ such terms,
1837: each having magnitude $1\over {N_A^2}$. Using the random phase approximation
1838: and including the sum over $j,k,b$ we get
1839: $ \approx {1\over N_A^2}N_A N_i \sqrt N_a = {1\over \sqrt N_a}$.
1840: This is as expected since the averaging over $N_a$ fine grained
1841: micro-states is expected to produce just such a suppression. Indeed,
1842: this was the motivation for introducing the $a$ 
1843: variables. This also agrees with v. Neumann's estimates
1844: on noting that his ${N_a\over S_a}$ is our $N_a$.
1845: 
1846: The time-dependent term is nevertheless very important
1847: for determining the
1848: {\em thermalization time}. Thus we have the result that to a very good
1849: accuracy (of order $1\over N_a^{1/2}$), the effective density
1850: matrix for large times
1851: is given by
1852: \be \label{4.39}
1853: \rho~~= {1\over N_i}\sum_i~|i\ran \lan i|~~=~~{1\over N_A}~~\sum_A~~|A\ran
1854: \lan A|
1855: \ee
1856: 
1857: It should be noted that the emergence of the microcanonical distribution is
1858: happening for
1859: non-trivial systems. This can be seen by examining 
1860: eqns (\ref{4.16}-\ref{4.17}) at $t=0$ where
1861:  there is no
1862: trace of the microcanonical distribution anywhere.
1863: 
1864: In section 6 we will calculate the leading order time dependent
1865: correction in a simple example that confirms the above expectations.
1866: 
1867: 
1868: One should also ask what happens if we start off the system in an exact
1869: energy eiegenstate. In that case we get for the expectation value
1870: of $O$
1871: \be \label{4.40}
1872: \lan \psi (t) | O \psi (t) \ran = \sum _i p_A ^i O_{ii} +
1873: \sum _{ij}R_{AA}^{ij}O_{ij}
1874: \ee
1875: 
1876: Note that if $O$ is a reasonable observable we should find
1877: that $\lan i| O^2|i\ran - \lan i|O|i\ran^2 < \lan i|O|i\ran ^2$.
1878: This means that $\sum _j O_{ij}O_{ji} \approx O_{ii}^2$. Thus
1879: one expects ${O_{ii}\over O_{ij}} \approx \sqrt N_j$ if we assume
1880: random phases or ${O_{ii}\over O_{ij}} \approx  N_j$ if they add in phase.
1881: If we now take into account this relative magnitudes
1882: of $O_{ij}$ versus $O_{ii}$, one gets that the off diagonal term is suppressed
1883: by a factor between $\sqrt{N_aN_i}$ and ${\sqrt {N_a}} N_i$. If $N_a$ is
1884: sufficiently large this could mean that for such systems even
1885: if the system starts of in an exact energy eigenstate it can be described
1886: by the microcanonical distribution. 
1887: 
1888: \section{Soft Quanta and the Resonance Picture}
1889: \setcounter{equation}{00}
1890: 
1891: In this section we will attempt to describe the general
1892: phenomenon of thermalisation in slightly different terms. Instead
1893: of working with the full Hamiltonian of the $i$ and $a$ system
1894: and the exact eigenstates, let us consider the same system from the point
1895: of view of time dependent perturbation theory. We think of the
1896: $i$ system as an atom with discrete states and the $a$ system as
1897: a collection of harmonic oscillators describing long wavelength
1898: photons.
1899: 
1900: As a first step let us treat the electro-magnetic field
1901: classically and let us think of the atom as a two level system. We are
1902: familiar with this ``NMR'' type of situation. When the electro-magnetic
1903: RF-pulse is the right frequency, we have resonant absorption
1904: and emission of photons. As we apply the RF pulse we expect the
1905: two level system to oscillate between the two states at the Bohr frequency
1906: and on the average there is equal probability to be in either
1907: of the two states. This happens even when the coupling is very weak.
1908: In terms of time independent perturbation theory (in the NMR case,
1909:  we can go to the rotating frame and make the problem
1910: time independent) the off diagonal terms are very large, close to
1911:  resonance,
1912: compared to the energy splitting. Thus the energy eigenstates have
1913: equal amounts of up and down spin states.
1914: 
1915: This physical intuition can be applied to our problem. First
1916: of all the electro-magnetic field has to be treated quantum mechanically.
1917: Thus there is spontaneous emission in addition to induced emission
1918: and absorption. If the strength of the electro-magnetic field
1919: is large then this can be neglected. In oscillator language,
1920: the field oscillators should be excited to a high enough quantum number.
1921: 
1922: Secondly, and more importantly we have a continuum of oscillators,
1923: both for emission and absorption. Here again we can apply our
1924: intuition from atomic physics. The effect of the continuum on the
1925: discrete state is to introduce an anti-Hermitian term in the effective
1926: Hamiltonian that gives the discrete states an imaginary ``width''.
1927: This makes the time evolution non-unitary and gives the states
1928: a lifetime by making the wave function decay. Similar mathematical
1929: manipulations can be done for the traced density matrix calculation to
1930: show within a perturbative approach that the off diagonal terms decay with time
1931: and the diagonal terms tend
1932: to become equal. The decay of the off diagonal terms was explained
1933: in Section 3 as being due to random phase cancelation. 
1934: Here we
1935: give arguments for seeing this more explicitly. Just as
1936: in the usual atomic physics calculation, the
1937: excited state decays by emission of energy in the form of a photon, similarly
1938: the off diagonal terms also decay as the photon is emitted and
1939: absorbed repeatedly. Just as the photon escapes
1940: in an apparently irreversible way into the continuum,
1941: never to return
1942: (in the limit of the Poincare recurrence time $\rightarrow \infty$),
1943:  so do the correlations escape irreversibly into the
1944: continuum. \footnote{In the AdS/CFT correspondence, this continuum
1945: of soft photons represents the far infrared of the CFT. In the
1946: bulk, this would be the region in the interior of the black hole.
1947: The irreversibility described above is the irreversibility of the black
1948: hole horizon. This was discussed in \cite{KRBS}}.
1949: 
1950: This language is also suitable for explaining intuitively why the
1951: traced matrix has equal diagonal terms. When the number
1952: density of soft quanta is large, then the probability,
1953: $P_{ij}$, of transition from $i$ to $j$ is equal to the reverse.
1954: We expect that the change in the $i$'th diagonal term in the
1955: density matrix is proportional to $P_j P_{ji}-P_i P_{ij}$.
1956: When $P_{ij}=P_{ji}$, this will stop changing only when
1957: $P_i=P_j$. This does not
1958: constitute a proof that the system will go toward equilibrium in
1959: general. But it allows us to intuitively understand a criterion
1960: for the emergence of microcanonical distribution. The main requirements are thus
1961: two:
1962: 
1963: i) There should be a continuum of soft excitations with energies
1964: containing the range of $\delta _i $ to $\Delta E$ in order for
1965: resonant absorption and emission to take place.
1966: In particular if $\delta _i > \Delta E$ there will not be
1967: any ergodization.
1968: 
1969: ii)The number of these should be large. We expect this number to be
1970: proportional to $E_{\gamma}\over {\hbar \omega}$, where $E_\gamma$
1971: is the total energy in soft quanta. Since $\hbar \omega \le \Delta E$,
1972: we have $E_\gamma >> \Delta E$. Also one expects $E >> E_\gamma$.
1973: Thus $E >> \Delta E$, as expected.
1974: 
1975: If we have a large number of
1976: hard photons as well, i.e. those with frequency larger than
1977: $\Delta E$, one would have to include them in the
1978: $i$-system. Otherwise the microcanonical description would not
1979: be a good approximation.
1980: 
1981: 
1982: While the discussion in Section 4 is completely general, it
1983: is somewhat abstract.
1984: The picture
1985: in this section is more intuitive. We can use this picture
1986: to attempt an answer to the question of when, in a given
1987: system, one can expect self thermalization. The difficult
1988: part presumably, is to identify the $i$ and $a$ variables
1989: unambiguously.
1990: \section{Two Discrete States Coupled to a Continuum}
1991: \setcounter{equation}{00}
1992: 
1993: The $|i\rangle \otimes |a\rangle$ system can be thought of as being composed
1994: of an atom coupled to the electro-magnetic field. This is denoted schematically
1995: in Figure 1 below. What is shown are the energy levels of the uncoupled
1996: system or equivalently, the situation where the coupling constant is
1997: set to zero. When the interaction is turned on we expect the energy
1998: levels of the atom to get broadened.
1999: 
2000: \begin{figure}[htbp]
2001: \begin{center}
2002: \epsfig{file=qsm.eps, width= 12 cm,angle=0}
2003: \vspace{ .2 in }
2004: \begin{caption}
2005: {Discrete states coupled to a continuum}
2006: \end{caption}
2007: \end{center}
2008: \label{fig2}
2009: \end{figure}
2010: 
2011: \subsection{Relaxation of Density Matrix - Thermalization}
2012: 
2013: We make the approximation namely that the $i$ system has only
2014: two states, $|0\rangle $ and $|1\rangle$. 
2015: Furthermore we will assume that the continuum degrees are those
2016: of a continuum
2017: of harmonic oscillators. Harmonic oscillators
2018: are easy to analyze and also representative of the real
2019: situation where we expect $a$ degrees to be soft quanta. 
2020: The frequencies
2021: of the harmonic oscillators are assumed to start at zero and form a continuum.
2022: In practice we just need the discrete spacing to be very small
2023: compared to the (inverse) time scales of interest. So if $\omega$ denotes 
2024: the frequency of the oscillators,
2025: ${1\over \delta E_a} = {1\over \delta \omega} >> T$. On the other hand
2026: $T$ should be large enough that the system has time to thermalize.
2027: In the previous section we saw that this time scale was set by $\gamma$.
2028: Thus $\gamma T > 1$ and so $\delta \omega << \gamma$. $\gamma$
2029: in turn is fixed by the matrix elements $\langle 1,0|V|0,a\rangle$.
2030: Thus here also we require that the off diagonal terms generated by
2031:  the interaction with the $a$ degrees of freedom must not be too small -
2032: this is expected whenever one has a resonance situation , i.e.
2033: $\hbar \omega = \Delta E$.
2034: 
2035: To do the calculation we will use the technique of influence
2036: functionals due to Feynman and Hibbs \cite{feyn}.
2037: 
2038: Following \cite{feyn} we let 
2039: $q$ be the coordinates of the $i$ system 
2040: \footnote{Though the $q$-system is actually a two level
2041: system, we have adopted a notation wherein the $q$-variables look continuous. The path integrals over $q$ are
2042: to be understood as appropriate matrix elements in the
2043: two level system.}
2044: and $Q$ that of the $a$ system. If the system starts
2045: off at time $t=0$ in the state $\Psi (q_i,Q_i,0)$, then the
2046: density matrix at time $t=T$ is given by (The subscript $i$ stands for
2047: `initial', and $f$ for `final')
2048: 
2049: \[
2050: \rho (q_f,q'_f,Q_f,Q'_f;T)=
2051: \int ...\int dq_i dq'_i dQ_i dQ'_i
2052: {\cal D} q(t){\cal D}q'(t) {\cal D}Q(t) {\cal D} Q'(t)
2053: \]
2054: \be
2055: e^{i[S(q)-S(q')] + i[S(Q)-S(Q')] +i[S_{int}(Q,q)-S_{int}(Q',q')]}
2056: \Psi (q_i, Q_i,0)\Psi ^* (q'_i,Q'_i,0)
2057: \ee
2058: 
2059: with boundary conditions:
2060: \[
2061: q(0)=q_i , q(T)=q_f , Q(0)=Q_i, Q(T)=Q_f
2062: \]
2063: \be
2064: q'(0)=q'_i , q'(T)=q'_f , Q'(0)=Q'_i, Q'(T)=Q'_f
2065: \ee
2066: The action $S(q)$ for the $i$-system will not be specified. However we will
2067: assume, as stated before that there are only two discrete states
2068: $|0\rangle , |1 \rangle$. We will further make the simplifying 
2069: assumption that $\langle 0 | q |0\rangle =\langle 1 | q |1\rangle =0$
2070: and that $\langle 0 | q |1\rangle \ne 0$.
2071: 
2072: The $a$-system is a harmonic oscillator. Actually it is an infinite number
2073: of harmonic oscillators, so in fact there should be an infinite number
2074: of $Q$ variables. For transparency of presentation 
2075: we first consider just one of them. As they are
2076: independent the generalization to arbitrary numbers
2077: or even a continuum is easy.
2078: 
2079: The wave function, $\Psi (q_i,Q_i,0)$, is assumed to factorize as
2080: \be
2081: \Psi (q_i,Q_i,0)= \phi _I (q_i)\chi _a (Q_i)
2082: \ee
2083: 
2084: Here
2085: \be
2086: |\phi _I\rangle = \alpha |0\rangle + \beta |1 \rangle
2087: \ee
2088: with  $ |\alpha |^2 + |\beta |^2 =1$. Let us first consider the case
2089: \be
2090: \chi (Q) \approx e^{-{Q^2\over 2 \omega}}
2091: \ee
2092: 
2093: which is the ground state of a harmonic oscillator of frequency $\omega$.  Furthermore
2094: we take $S_{int}(q,Q) \approx C\int dt q(t)Q(t)$.
2095: 
2096: We are interested in the density matrix traced over the $Q$ variables:
2097: \[
2098: \int dQ_f \rho (q_f,q'_f,Q_f,Q_f,T) \equiv \rho _Q(q_f,q_f',T)
2099: \]
2100: If we imagine doing all the $Q,Q'$ integrals we get an expression of the form
2101: \be   \label{rho}
2102: \rho _Q(q_f,q'_f,T)= \int ...\int dq_i dq'_i {\cal D} q(t){\cal D}q'(t)
2103: e^{i[S(q)-S(q')]}\phi _I(q_i)\phi ^*_I(q'_i)F(q,q').
2104: \ee
2105: 
2106: Here $F(q,q')$ is the result of doing all the $Q$ integrals and incorporates
2107: the entire effect of the $Q$-system, called 
2108: {\em environment} by Feynman and Hibbs, on the 
2109: $q$-system. This is the influence functional.
2110: For the special case of quadratic $S(Q)$ and bilinear $S_{int}(q,Q)$ it has been
2111: shown in \cite{feyn} that it has the form,
2112: \be   \label{IF1}
2113: F(q,q')=e^{-\int \int dt dt' [q(t)-q'(t)][\alpha (t,t')q(t')-\alpha ^* (t,t')q'(t')]}
2114: \ee
2115: 
2116: For a harmonic oscillator of frequency $\omega $,
2117: in its {\em ground state}
2118: we have $\alpha (t,t')= \alpha (t-t')$ and
2119: also
2120: \be   \label{alpha}
2121: \int d\tau e^{-i\nu \tau }\alpha (\tau )
2122: \equiv a(\nu ) \approx {C\over \omega ^2}\delta (\nu +\omega )
2123: \ee
2124: This is what Feynman and Hibbs call a
2125: {\it cold-environment} and for such a situations
2126: transitions pumping energy into the system are unlikely.
2127: 
2128: When we have a continuum of oscillators we simply have to integrate this
2129: expression (\ref{alpha}) over $\omega$.
2130: We thus insert (\ref{IF1}) into (\ref{rho}) and evaluate it perturbatively.  
2131: 
2132: \subsubsection{Zeroth Order}
2133: 
2134: \br
2135: \rho _{00}= \int dq_f dq_f' \phi _0^*(q_f)\phi _0 (q_f')\rho_Q(q_f,q'_f,T)\nonumber \\
2136: \rho _{10}= \int dq_f dq_f' \phi _1^*(q_f)\phi _0 (q_f')\rho_Q(q_f,q'_f,T)
2137: \er
2138: and so on.
2139: Thus
2140: \be
2141: \lan 0 |\rho |0\ran = 
2142: \lan 0 | e^{-iHT}(\alpha |0\ran + \beta |1\ran )(\lan 0|\alpha ^* + \lan 1|\beta ^* )
2143: e^{iHT}|0\ran = |\alpha |^2
2144: \ee
2145: \be
2146: \lan 1 |\rho |1\ran = |\beta |^2
2147: \ee
2148: \be
2149: \lan 1 |\rho |0\ran = 
2150: \lan 1 | e^{-iHT}(\alpha |0\ran + \beta |1\ran )(\lan 0|\alpha ^* + \lan 1|=
2151: \beta ^* )
2152: e^{iHT}|0\ran = \alpha ^* \beta e^{-i(E_1-E_0)T}
2153: \ee
2154: \be
2155: \lan 0 |\rho |1\ran = \alpha \beta ^* e^{+i(E_1-E_0)T}
2156: \ee
2157: 
2158: 
2159: \subsubsection{First Order}
2160:  We now go to the next order. We bring down one power of $\alpha (t,t')$.
2161: 
2162: We write it as the sum of four terms:
2163: \[
2164: \delta \rho _{Qj0}(T)=
2165: -\int \int dq_f dq'_f \phi ^* _j(q_f)\phi _0 (q_f') \int {\cal D}q(t){\cal =
2166: D}q'(t)
2167: \int dq_i \int dq'_i
2168: \]
2169: \be   \label{drho}
2170: e^{i[S(q)-S(q')]}[a+b+c+d]\phi _I (q_i)\phi _I^*(q'_i)
2171: \ee
2172: 
2173: where the subscript $j$ on $\phi$ can be  $0,1$ and
2174: \br
2175: a &=& \int dt \int dt' q(t)q(t')\alpha (t,t') \nonumber \\
2176: b &=& -\int dt \int dt' q'(t)q(t')\alpha (t,t') \nonumber \\
2177: c &=& -\int dt \int dt' q(t)q'(t')\alpha ^*(t,t') \nonumber \\
2178: d &=& \int dt \int dt' q'(t)q(t')\alpha ^*(t,t')
2179: \er
2180: 
2181: The calculations are done in the appendix. Here we will simply give the result.
2182: 
2183: We get for the first order correction to $\rho _{00}$ the following:
2184: \be   \label{31}
2185: \delta \rho _{Q00}(T)= -T\{ |\alpha |^2 [2a_R(\Delta E) + 2 a_R(-\Delta E)]-
2186: 2a_R(-\Delta E) \}|\lan 0 |q|1\ran |^2
2187: \ee
2188:  where $\Delta E = E_1-E_0$, and
2189: 
2190:  Let us first assume that $\Delta E >0$. Using (\ref{alpha}), which says that
2191: $a_R(\Delta E )=0$, we find
2192: \be   \label{32}
2193: \delta \rho_{Q00}(T) = -T[(|\alpha |^2 -1)2a_R(-\Delta E )]|\lan 0|q|1\ran |^2
2194: \ee
2195: This goes to zero when $|\alpha | \rightarrow 1$.
2196: 
2197: If $\Delta E <0$ we get
2198: \be   \label{33}
2199: \delta \rho _{Q00}(T)= -T |\alpha |^2 2 a_R(\Delta E )|\lan 0|q|1\ran |^2
2200: \ee
2201: 
2202: When $E_1 > E_0$ one expects the system to decay to its lower energy state
2203: $|0\ran$ with the emission of a soft quanta. The decay stops when the
2204: probability of occupation of the lower energy state is one. Also as (\ref{32})
2205: shows $\delta \rho$ is increasing with time as it should be.
2206: Thus one expects $\beta (T) \approx \beta (0) e^{-\gamma T}$.  So
2207: $|\alpha (T)|^2 = 1- |\beta (T)|^2 \approx |\alpha (0)| ^2 +|\beta (0)|^2
2208: 2\gamma T$.  Thus $\delta \rho \approx |\beta (0)|^2 2\gamma T$. This agrees
2209: with (\ref{32}) for $\gamma = a_R(-\Delta E)|\lan 0|q|1\ran |^2$.
2210: 
2211: 
2212: 
2213: Similarly when $E_0>E_1$, the probability of occupation of $E_0$
2214: decreases with time as $|\alpha (T)|^2 \approx |\alpha (0)|^2 e^{-2\gamma T}
2215: $ and so $\delta \rho \approx -|\alpha (0)|^2 2\gamma T$ as
2216: shown by (\ref{33}).
2217: 
2218: 
2219: Similarly we can look at the off diagonal terms
2220: \[
2221: \delta \rho_{Q10} = -T \alpha ^* \beta e^{-i(E_1-E_0)T}|\lan 0|q|1\ran |^2
2222: [a(-\Delta E) + a(\Delta E)^*]+
2223: \]
2224: \be   \label{34}
2225: [\int _0^Td\tau e^{-i(E_1-E_0)\tau}] \alpha \beta ^* e^{-i(E_1-E_0)T}
2226: (\lan 1 |q |0\ran )^2 [a(0) + a(0)^*]
2227: \ee
2228: 
2229: 
2230: If $\Delta E \ne 0$, the integral over $\tau$ in the second term of
2231: (\ref{34}) is vanishingly small relative to the first term. We are left with
2232: \be \label{6.20}
2233: \delta {\rho}_{Q10} = -T \alpha ^*\beta e^{-i(E_1-E_0)T}|\lan 0|q|1\ran |^2
2234: [a(\Delta E)^* + a(-\Delta E )]
2235: \ee
2236: This is also as expected because the product $\alpha (T) \beta (T)$
2237: decays as $e^{-\gamma T}$ (one of the terms decays and the other goes
2238: to 1).
2239: 
2240: 
2241: While all this is
2242: as expected, this does not really demonstrate the De-coherence that we
2243: are after. In order to demonstrate decoherence we must show that
2244: the off diagonal terms decay even when the diagonal terms do not.
2245: In order to obtain a situation where the probabilities of occupation
2246: of the states $|0\ran ,|1\ran$, are equal, one must have the external
2247: oscillators in excited states as well so that the system can absorb soft
2248: quanta.
2249: An easy route would be to let the two states be exactly
2250: degenerate so that energy conservation allows transitions
2251: in either direction. But in such a situation it is always
2252: possible to find orthonormal linear combinations of the two
2253: states in which the the operator $q$ is {\it diagonal}. It is
2254: then a simple exercise to show that in this basis the diagonal
2255: elements of the density matrix do not change at all. Even on
2256: physical grounds exact degeneracy is unreasonable to assume.
2257: What is more reasonable is to consider a situation
2258: in which $\Delta E\ne 0$ but "small" compared to some
2259: relevant energy scale
2260: in the problem. Nevertheless, we need to calculate influence
2261: functionals for situations where the system can both absorb
2262: and give energy to the quantum environment. The influence
2263: functional calculated in Feynman and Hibbs for the case where
2264: all the environment oscillators are in the ground state will
2265: not suffice. While the generic case of the environment
2266: oscillators in arbitrary excited states is too difficult to
2267: handle, the influence function for the case where the
2268: environment states are {\it coherent states} or {\it number states}
2269: can be
2270: explicitly evaluated and this suffices for our purpose.
2271: 
2272: A partial form of this result is already available in Feynman
2273: and Hibbs (see eqn 8.141). They considered the amplitude
2274: $f(b,a)$
2275: that a {\it forced} harmonic oscillator goes from state
2276: $f$ to state $g$ where
2277: \be
2278: f(x)~=~({M\omega\over \pi\hbar})^{1\over 4}
2279: ~e^{-({M\omega\over \pi\hbar})(x-a)^2}~~~~~~~~
2280: g(x)~=~({M\omega\over \pi\hbar})^{1\over 4}
2281: ~e^{-({M\omega\over \pi\hbar})(x-b)^2}
2282: \ee
2283: Their explicit expression for $f(b,a)$ is
2284: \be
2285: f(b,a)~=~A~e^{[-{M\omega\over 4\hbar}(a^2+b^2-2ab~e^{-i\omega T})+
2286: i({M\omega\over 2\hbar})^{1\over 2}
2287: (a\eta+b\eta^*e^{-i\omega T})]}
2288: \ee
2289: where
2290: \be
2291: A~=~e^{-i\omega T/2}
2292: ~e^{
2293: -{1\over 2M\omega\hbar} \int_0^T~\int_0^t~\gamma(t)\gamma(s)
2294: e^{-i\omega(t-s)}
2295: }
2296: \ee
2297: Of course, the states $f(x),g(x)$ are {\em coherent states}
2298: of the harmonic oscillator but with {\em real} parameters.
2299: It suffices, for our purposes, to generalise this result to the case where $f(x)$ is a coherent state with {\it complex}
2300: parameter $\xi$ while leaving $b$ still real. The
2301: corresponding results  are
2302: \be
2303: f(x)~=~({M\omega\over \pi\hbar})^{1\over 4}
2304: ~e^{{\xi^2\over 2}-{|\xi|^2\over 2}}
2305: ~e^{-({M\omega\over \pi\hbar})(x-\sqrt{{2\hbar\over M\omega}}~\xi)^2}~~~~~~~~
2306: \ee
2307: where we have followed the Pancharatnam convention for phases.
2308: \be
2309: f(b,\xi)~=~{\cal A}~e^{[-{M\omega\over 4\hbar}(
2310: {2\hbar\over M\omega}
2311: \xi^2+b^2-2
2312: \sqrt{{2\hbar\over M\omega}}
2313: \xi b~e^{-i\omega T})+
2314: i({M\omega\over 2\hbar})^{1\over 2}
2315: (
2316: \sqrt{{2\hbar\over M\omega}}
2317: \xi\eta+b\eta^*e^{-i\omega T})]}
2318: \ee
2319: where
2320: \be
2321: {\cal A}~=~e^{{\xi^2\over 2}-{|\xi|^2\over 2}}
2322: ~e^{-i\omega T/2}
2323: ~e^{
2324: -{1\over 2M\omega\hbar} \int_0^T~\int_0^t~\gamma(t)\gamma(s)
2325: e^{-i\omega(t-s)}
2326: }
2327: \ee
2328: Now it is easily seen that $G_{0\xi},G_{m\xi}$
2329: where
2330: $G_{fi}$ is the amplitude that the oscillator initially
2331: in state $i$ is found in state $f$ at time T, are given
2332: by
2333: \be
2334: G_{0\xi}=~e^{-{|\xi|^2\over 2}+i\xi\eta}
2335: ~e^{
2336: -{1\over 2M\omega\hbar} \int_0^T~\int_0^t~\gamma(t)\gamma(s)
2337: e^{-i\omega(t-s)}
2338: }
2339: \ee
2340: and
2341: \be
2342: G_{m\xi}=~{1\over {\sqrt{m!}}}~G_{0\xi}~(\xi+i\eta^*)^m
2343: \ee
2344: Finally the influence functional is given by
2345: \be
2346: F[q,q^{\prime}]~=~\sum_m~G_{m\xi}{G^{\prime}_{m\xi}}^*
2347: \ee
2348: The final result can be expressed as
2349: \be
2350: F_{\xi}[q,q^{\prime}]~~e^{i\xi(\eta-\eta^{\prime})+i\xi^*(\eta^*-{\eta^{\prime}}^*)}
2351: ~F_0[q,q^{\prime}]
2352: \ee
2353: where $F_0$ is the influence functional for the case
2354: where the environment oscillators are in the ground
2355: state.
2356: 
2357: Thus the environmental coherent states produce an
2358: additional influence functional whose exponent is a
2359: linear functional of $q$. However, to the degree of
2360: accuracy needed for our perturbative calculations
2361: we can translate this into an effective quadratic
2362: functional by expanding the additional term to second
2363: order i.e
2364: $$
2365: -{1\over 4M^2\omega}[\xi^2~(\eta-\eta^{\prime})^2
2366: +{\xi^*}^2~(\eta^*-{\eta^{\prime}}^8)^2
2367: +2|\xi|^2~|(\eta-\eta^{\prime})|^2]
2368: $$
2369: It is sufficient to look at terms of the type
2370: $$
2371: \int_0^T~dt\int_0^t~ds \Delta a(t,s) q(t) q(s)
2372: $$
2373: to determine the {\em one complex function} $\Delta a(t,s)$
2374: that determines the additional effective influence functional.
2375: After some algebra it follows that
2376: \be
2377: \Delta a(t,s)~=~{C^2\over 4M^2\omega}~
2378: [2\xi^2 e^{-i\omega(t+s)}+2{\xi^*}^2 e^{i\omega(t+s)}+
2379: 2|\xi|^2 (e^{i\omega (t-s)}+e^{-i\omega(t-s)}]
2380: \ee
2381: Unlike the case of the environment oscillators in the
2382: ground state the complex function is not a function of
2383: $t-s$ alone but it turns out that the parts of $a(t,s)$
2384: that are functions of $t+s$ do not contribute to the
2385: transition rates as the integrations over $s,t$ produce
2386: mutually exclusive delta functions and one is
2387: effectively left with
2388: \be \label{6.32}
2389: \Delta a_{eff}(t,s)~=~{C^2\over 4M^2\omega}~
2390: 2|\xi|^2 [(e^{i\omega (t-s)}+e^{-i\omega(t-s)}]
2391: \ee
2392: Thus when the environmental oscillators are in  coherent
2393: states we have a mixture of a {\it cold-environment}
2394: whose strength is {\em independent} of the coherent
2395: state parameter $\xi$ and a {\it classical noise}
2396: environment with $a_R(\nu)=a_R(-\nu)$. Further $a_R(\nu)$ is
2397:  directly proportional to the level of excitation
2398: of the coherent states. It should be
2399: emphasised that any environment can be represented
2400: as a mixture of a cold environment and a classical
2401: noise environment. But what is special about the
2402: coherent state case is that the cold component is
2403: {\em independent} of the level of excitation of the coherent states.
2404: This will be crucial for us later on.
2405: 
2406: In the number state case, there are no $t+s$ terms to begin
2407: with (as shown on the Appendix) and one obtains (\ref{C28})
2408: directly - this is (\ref{6.32}) with $|\xi |^2 $
2409: replaced by $m$, the excitation number of the oscillator.
2410: 
2411: If we use (\ref{C28}) (or (\ref{6.32})), with large $m$
2412: or $|\xi |^2$, we can neglect the cold term in the
2413: influence functional. In this case $a_R(\Delta E)=
2414: a_R (-\Delta E)$, and both are non-zero. Then we
2415: see from (\ref{6.20}) that
2416: \[
2417: \delta \rho _{Q00} = -T (2 |\alpha | ^2 -1) |\lan 0|q|1\ran|^2 2a_R(\Delta E)
2418: \]
2419: \[
2420: \delta \rho _{Q01}= -T \underbrace
2421: {\alpha ^* \beta e^{-i\Delta E T}}_{\rho _{Q01}}|\lan 0 |q|1\ran|^2
2422: 2a_R(\Delta E)
2423: \]
2424: \be \label{6.33}
2425: \approx ~\rho _{Q01}(e^{-\gamma T}-1)
2426: \ee
2427: Thus in this approximation
2428: even when $|\alpha |^2={1\over2}$, $\delta \rho _{Q01}\ne 0$.
2429: Equation (\ref{6.33}), if our extrapolation from the linear term
2430: to the full exponential is correct, demonstrates decoherence. To actually
2431:  prove decoherence
2432: one would have get a more complete solution. Presumably this can be done numerically
2433: in some cases.
2434: 
2435: 
2436: Thus we have evidence for  decoherence in an explicit calculation in
2437: this simplified model. The crucial point here is the appearance
2438: of the imaginary term due to ``absorption of a soft quanta by the
2439: environment''. We have put the quotation marks because this is really
2440: just a convenient trick. We have made a division into two sets of variables
2441: $i$ and $a$ with the requirement that $a$ should have a continuous spectrum
2442: We called these ``soft quanta''. Then we showed that the same mathematical
2443: manipulations that give rise to an imaginary part to the energy of a
2444: discrete state can be used here to show that the off-diagonal terms of the 
2445: density matrix decay.
2446: 
2447: We should also check that the conditions set forth at the end
2448: of the last section are indeed satisfied. $\delta _a$ is zero for this
2449: calculation.
2450: Due to resonance, the off-diagonal terms are large.
2451: Thus both conditions
2452: are satisfied. While we
2453: have not {\em proved} even in a non-rigourous way,
2454: the existence of the phenomenon of self thermalization,
2455: the calculations done here do make plausible the physical ideas of section
2456: 2 and also buttress the intuitive random phase arguments
2457: about the off diagonal terms of the density matrix
2458: decaying in time, that were made in section 3.
2459: 
2460: \section{Conclusions}
2461: 
2462: In this paper we have attempted to provide a physical picture
2463: of the process of self thermalization, by which a pure state
2464: can appear thermal if studied with coarse resolution. This phenomenon
2465: underlies the validity of the microcanonical ensemble in quantum statistical
2466: mechanics. It also has applications in the black hole information
2467: paradox.
2468: 
2469: We have presented an intuitive
2470: physical picture along with a mathematical formulation
2471: of the process. The principal new ingredient was the introduction
2472: of dynamical ``soft quanta'' that are not included in the
2473: microstate descriptions. They are
2474: coarse grained away and
2475: in the process, remove correlations between the (coarse grained) microstates.
2476: Thus they not only provide the quantum analogue of the Gibbsian coarse
2477: graining of microstates (Section 4), they also thermalize the system
2478: by resonant absorpton and emission (Section 5,6). Mathematically
2479: this is identical to the appaearance of an imaginary part in the
2480: energy of an excited state when it couples to a continuum.
2481: This also provides a  way to calculate the thermalization rate.
2482: A first order perturbation calculation (Section 6) supports this picture.  
2483: 
2484: It would be interesting to apply this picture to some interesting
2485: systems in a quantitative way. This would be a test of the correctness
2486: of these ideas.
2487: 
2488: \noindent
2489: {\bf Acknowledgements}\\ We would like to thank Balram
2490: Rai and
2491: R. Shankar for discussions.
2492: 
2493: 
2494: 
2495: \appendix
2496: 
2497: \renewcommand{\thesection}{\Alph{section}}
2498: %\renewcommand{\thesubsection}{{\subsection}}
2499: \renewcommand{\theequation}{\thesection.\arabic{equation}}
2500: 
2501: \section{Appendix: Density Matrix Calculation}
2502: \label{appena}
2503: \setcounter{equation}{0}
2504: In this appendix we derive (\ref{31}) and (\ref{34}) starting from (\ref{drho}).
2505: 
2506: 
2507: We will first work out in detail the contribution of the perturbation labelled
2508: `$a$'.
2509: 
2510: The $q$ and $q'$ integrals can be done separately. Let us denote by
2511: $A_j$ the $q$ integral and by $A'$ ($A''$ does not depend on $j$) the $q'$ integral.
2512: Thus
2513: \[
2514: A_j=\int dq_f \phi ^*_j (q_f){\cal D}q e^{iS(q)}\int dq_iq(t)q(t')\phi _I(q_i)
2515: \]
2516: \[
2517: A'= \int dq'_f \phi _0(q'_f){\cal D}q'(t) e^{-iS(q')}dq'_i \phi _I^*(q'_i)
2518: \]
2519: \be
2520: \delta \rho _{Qj0}(T) = \int _0^T dt\int _0^T dt' \alpha (t,t') A_j A'
2521: \ee 
2522: 
2523:  
2524: {$\bf A_j:$}
2525: 
2526: \[
2527: A_j=\int dq_f \int dq_i \phi ^* _j (q_f)\int {\cal D}q e^{iS(q)}q(t)q(t')
2528: \phi _I(q_i)
2529: \]
2530: \[
2531:  =
2532: \sum _{m=0,1}
2533: \lan j|e^{-iH(T-t)}q|m\ran e^{-iH(t-t')}\lan m|qe^{-iHt'}
2534: (\alpha |0\ran +\beta |1\ran )  
2535: \]
2536: 
2537: If $j=0$ we get
2538: \[
2539: A_0 =
2540: \lan 0 |e^{-i(E_0(T-t)} q | 1 \ran e^{-iE_1(t-t')}\lan 1 | qe^{-iE_0t'} \alpha |0\ran
2541: \]
2542: \be
2543: =e^{-iE_0T}e^{-i(E_1-E_0)(t-t')}\lan 0 |q|1\ran \lan 1 | q | 0 \ran \alpha
2544: \ee
2545: 
2546: If $j=1$ we get
2547: 
2548: \[
2549: A_1 = \lan 1 | e^{-iE_1(T-t)}q | 0\ran e^{-iE_0(t-t')}\lan 0| qe^{-iE_1t'}\beta |1\ran
2550: \]
2551: \be
2552: = e^{-iE_1T}e^{-i(E_0-E_1)(t-t')}\lan 1|q |0 \ran \lan 0 | q | 1 \ran \beta
2553: \ee
2554: 
2555: {$\bf A':$}
2556: 
2557: \[
2558: A'=[ \int dq'_f {\cal D} q'(t) dq'_i \phi _I (q'_i)e^{iS(q')}\phi _i^*(q'_f)]^*
2559: \]
2560: \be
2561: =
2562: [\lan 0 | e^{-iHT}(\alpha | 0 \ran + \beta | 1 \ran ) ] ^*
2563: =\alpha ^* e^{iE_0T}
2564: \ee
2565: 
2566: 
2567: Combining all the above we get the contribution due to the perturbation
2568: $a$.
2569: :  
2570: 
2571: \[
2572: \delta \rho _{Q00}(T)= |\alpha |^2 \int _0^Tdt \int _0^T
2573: dt' e^{-i(E_1-E_0)(t-t')}\alpha (t-t')
2574: |\lan 1 |q|0\ran |^2
2575: \]
2576: \be
2577: = T |\alpha |^2 |\lan 1 |q|0\ran |^2 a(E_1-E_0)
2578: \ee
2579: 
2580: and
2581: 
2582: 
2583: \[
2584: \delta \rho _{Q10}(T) = \alpha ^* \beta e^{-i(E_1-E_0)T}\int _0^T dt
2585: \int _0^T dt'
2586: e^{-i(E_0-E_1)(t-t')} \alpha (t-t') |\lan 1 |q|0\ran |^2
2587: \]
2588: \be
2589: =
2590: T \alpha ^* \beta |\lan 1 |q|0\ran |^2 a(E_0-E_1)e^{-i(E_1-E_0)T}
2591: \ee
2592: 
2593: We can similarly calculate the contribution from the other ones.
2594: 
2595: 
2596: \noindent
2597: {$\bf B_j:$}
2598: 
2599: \[
2600: B_0= \lan 0|e^{-iH(T-t')}q|m\ran e^{-iE_mt'}\lan m
2601:  |(\alpha |0\ran + \beta |1\ran )
2602: \]
2603: \be
2604: = \beta e^{-iE_0T}e^{-i(E_1-E_0)t'}\lan 0 | q| 1 \ran
2605: \ee
2606: 
2607: 
2608: \[
2609: B_1=\lan 1|e^{-iH(T-t')}q|m\ran e^{-iE_mt'}\lan m | (\alpha
2610: | 0 \ran + \beta | 1 \ran )
2611: \]
2612: \be
2613: = \alpha e^{-iE_1T}e^{-i(E_0-E_1)t'}\lan 1 | q|0 \ran
2614: \ee
2615: $\bf B':$
2616: 
2617: \[
2618: B'=[ \int dq'_f \phi _i^*(q'_f){\cal D} q'(t) dq'_i e^{iS(q')}\phi _I (q'_i)]^*
2619: \]
2620: \[
2621: = [\lan 0 | e^{-iH(T-t)}q'|m\ran e^{-iE_m t}\lan m|
2622: (\alpha |0 \ran + \beta | 1 \ran )]^* 
2623: \]
2624: \be
2625: = \beta ^* e^{+iE_0T}\lan 1 | q | 0 \ran e^{-i(E_0-E_1)t}
2626: \ee
2627: 
2628: Combining the above
2629: 
2630: \be
2631: \delta \rho _{Q00}= T|\beta |^2 |\lan 0 |q|1\ran |^2 a(E_0-E_1)
2632: \ee
2633: and
2634: \be
2635: \delta \rho _{Q10}=[\int _0^T d\tau
2636: e^{-i(E_0-E_1)\tau}]
2637: \alpha \beta ^*e^{-i(E_1-E_0)T}(\lan 1 | q | 0 \ran )^2  a(0)
2638: \ee
2639: 
2640: (Note that the factor inside the square brackets replaces the factor
2641: $T$ that appeared in previous terms.)
2642: 
2643: 
2644: $\bf C_j:$
2645: 
2646: \[
2647: C_0 = \lan 0| e^{-iH(T-t)}q|m\ran e^{-iE_mt}\lan m | (\alpha |0\ran + \beta
2648: |1 \ran )
2649: \]
2650: \be
2651: = \beta e^{-iE_0 T}e^{-i(E_1-E_0)t}\lan 0 | q | 1\ran
2652: \ee
2653: 
2654: \[
2655: C_1= \lan 1 | e^{-iH(T-t)}q|m\ran e^{-iE_mt}\lan m | (\alpha |0\ran + \beta
2656: |1 \ran )
2657: \]
2658: \be
2659: =\alpha e^{-iE_1 T}e^{-i(E_0-E_1)t}\lan 1| q | 0\ran
2660: \ee
2661: 
2662: $\bf C':$
2663: 
2664: \[
2665: C'= [ \int dq'_f \phi _i^*(q'_f){\cal D} q'(t') dq'_i e^{iS(q')}
2666: \phi _I (q'_i)]^*
2667: \]
2668: \[
2669: = [\lan 0 | e^{-iH(T-t')}q'|m\ran e^{-iE_m t'}\lan m|
2670: (\alpha |0 \ran + \beta | 1 \ran )]^* 
2671: \]
2672: \be
2673: =\beta ^*  e^{+iE_0 T}e^{-i(E_0-E_1)t'}\lan 1| q | 0\ran
2674: \ee
2675: 
2676: Combining:
2677: \be
2678: \delta \rho _{Q00}= T |\beta |^2  |\lan 0 | q | 1 \ran |^2 [a(E_0-E_1)]^*
2679: \ee
2680: 
2681: \be
2682: \delta \rho _{Q10}=[\int _0^T d\tau
2683: e^{-i(E_0-E_1)\tau}]
2684: \alpha \beta ^*e^{-i(E_1-E_0)T}(\lan 1 | q | 0 \ran )^2  a^*(0)
2685: \ee
2686: 
2687: 
2688: $\bf D_j:$
2689: 
2690: \be
2691: D_0= \lan 0|e^{-iHT}(\alpha |0\ran + \beta
2692: |1 \ran )
2693: =\alpha e^{-iE_0T}
2694: \ee
2695: \be
2696: D_1= \lan 1|e^{-iHT}(\alpha |0\ran + \beta
2697: |1 \ran )
2698: =\beta e^{-iE_1T}
2699: \ee
2700: 
2701: 
2702: 
2703: $\bf D':$
2704: \[
2705: D'=
2706: [ \int dq'_f \phi _i^*(q'_f){\cal D}q'(t) q'(t') dq'_i e^{iS(q')}
2707: \phi _I (q'_i)]^*
2708: \]
2709: \[
2710: = [\lan 0 | e^{-iH(T-t)}q'|m\ran e^{-iE_m (t-t')}\lan m|q e^{-iHt'}
2711: (\alpha |0 \ran + \beta | 1 \ran )]^* 
2712: \]
2713: \be
2714: = \alpha ^* e^{iE_0T}e^{-i(E_0-E_1)(t-t')}|\lan 0|q |1\ran |^2
2715: \ee
2716: 
2717: 
2718: Combining:
2719: \be
2720: \delta \rho _{Q00}= T |\alpha |^2  |\lan 0 | q | 1 \ran |^2 [a(E_1-E_0)]^*
2721: \ee
2722: 
2723: \be
2724: \delta \rho_{Q10}=
2725: T \alpha ^* \beta |\lan 1 |q|0\ran |^2 [a(E_1-E_0)]^*e^{-i(E_1-E_0)T}
2726: \ee
2727: 
2728: 
2729: If we add up all the contributions
2730: $a+b+c+d$ we get the expressions (\ref{31}) and
2731: (\ref{34}).
2732: 
2733: \section{Appendix: Influence Functional Calculation}
2734: \label{appenb}
2735: \setcounter{equation}{0}
2736: 
2737: In this Appendix we will derive the influence functional for the
2738: cases when the "external" oscillators are either in coherent
2739: states or in  number eigenstates.
2740: 
2741: We start with the kernel $K(Q_f,Q_i,Cq)$ for a forced
2742: harmonic oscillator forced with $f(t)=Cq(t)$ given by
2743: \be
2744: K= \underbrace {{\sqrt {M\omega \over 2\pi i \hbar sin ~ \omega T }}}_N e^{{i\over\hbar}
2745: S_{cl}(Q_f,Q_i,Cq)}
2746: \ee
2747: As shown in \cite{feyn} $S_{cl}$ is given by,
2748: \br
2749: S_{cl}&=& {M\omega\over 2 sin ~ \omega T}[cos ~\omega T (Q_f^2+Q_i^2)
2750: -2Q_fQ_i\nonumber\\
2751: &+&
2752: {2Q_fC\over M\omega} \int _0^T q(t) sin ~ \omega t ~dt +
2753: {2Q_iC\over M\omega} \int _0^T q(t)sin ~ \omega (T-t)~dt\nonumber\\
2754: &-&{C^2\over M^2\omega^2} \int _0^T\int _0^t q(t)q(s)sin ~\omega (T-t) 
2755: sin~ \omega s ~ds ~dt ]
2756: \er
2757: 
2758: We need to calculate the following quantity:
2759: \be \label{IF}
2760: \int dQ_f dQ_i dQ'_i |N|^2 e^{{i\over\hbar}[S_{cl}(Q_f,Q_i,Cq) -S_{cl}(Q_f,Q'_i,Cq')}
2761: \psi (Q_i) \psi ^* (Q_i')
2762: \ee
2763: The $Q$ oscillator starts off at time $t=0$ in a state $\psi$. The density matrix at time $T$ is then traced over the coordinates of
2764: the $Q$ oscillator.  For $\psi (Q_i)$ we use the coherent states.
2765: Thus we use
2766: \be
2767: f(x)~=~({M\omega\over 2\hbar})^{1\over 4}
2768: ~e^{{\xi^2\over 2}-{|\xi|^2\over 2}}
2769: ~e^{-({M\omega\over 2\hbar})(x-\sqrt{{2\hbar\over M\omega}}~\xi)^2}~~~~~~~~
2770: \ee
2771: where we have followed the Pancharatnam convention for phases.
2772: It should be noted that eqn (B.3) straightaway gives the
2773: influence functional. The method followed in \cite{feyn} is
2774: needlessly circuitous. After obtaining the {\em nonperturbative}
2775: expression for the coherent case influence functional which
2776: will be shown to have a part {\it linear} in $q(t),q^{\prime}(t)$,
2777: we will introduce an effective influence functional quadratic
2778: in $q(t),q^{\prime}(t)$ that is obtained perturbatively. We will
2779: obtain expressions for this effective influence functional
2780: both for the {\em coherent states} as well as for the {\em number states}.
2781: 
2782: The $Q$-dependent terms in the integrand in (\ref{IF}) can be written as
2783: \be
2784: e^{-{1\over 2} [A Q_f^2 + BQ_i^2 + C{Q'}_i^2 +2DQ_fQ_i +2EQ_fQ'_i ]
2785: + FQ_f + GQ_i + HQ'_i }
2786: \ee
2787: with
2788: \be
2789: A=0, ~~B=-{M\omega\over \hbar \sin {\omega T}}ie^{i\omega T}, ~~C=B^*,
2790: \ee
2791: \be
2792: D= {iM\omega\over \hbar\sin {\omega T}}~~,~~
2793: E=D^*
2794: \ee
2795: \be
2796: F= {iC\over \hbar\sin { \omega T}} \int _0^T (q-q') \sin {\omega t} ~dt,
2797: \ee
2798: \be
2799: G= {iC\over \hbar\sin { \omega T}} \int _0^T q sin ~ \omega (T-t) ~
2800: dt~+~\sqrt{{2M\omega\over\hbar}}\xi
2801: \ee
2802: \be
2803: H= {-iC\over \hbar\sin { \omega T}} \int _0^T q^{\prime} sin ~ \omega (T-t) ~
2804: dt~+~\sqrt{{2M\omega\over\hbar}}\xi^*
2805: \ee
2806: 
2807: The result of doing the Gaussian integral is
2808: \be
2809: \sqrt{{(2\pi)^3\over 2}({\hbar\over M\omega})^3 sin^2 ~\omega T}~~
2810: e^{{\hbar\over 4M\omega}[F^2 + (G+H)^2 + e^{-i\omega T}2FG + e^{i\omega T}2FH]}
2811: \ee
2812: The overall normalisation factor is
2813: \br
2814: {\cal N}&=&|N|^2~
2815: \sqrt{{(2\pi)^3\over 2}({\hbar\over M\omega})^3 sin^2 ~\omega T}~~
2816: ({M\omega\over \pi\hbar})^{1/2}~
2817: e^{\xi^2/2+{\xi^*}^2/2-|\xi|^2}\nonumber\\
2818: &=&
2819: e^{\xi^2/2+{\xi^*}^2/2-|\xi|^2}
2820: \er
2821:  Thus the final result is
2822: \br
2823: &~&{\cal N} exp \{{\hbar\over 4M\omega}[F^2 + (G+H)^2 + e^{-i\omega T}2FG +
2824:  e^{i\omega T}2FH] \nonumber\\
2825: &-&{i\over \hbar\sin{\omega T}}{C^2\over M\omega} \int _0^T\int _0^t q(t)q(s)sin ~\omega (T-t)
2826: sin ~\omega s ~ds ~dt \nonumber\\
2827: &+&{i\over \hbar\sin{\omega T}}{C^2\over M\omega} \int _0^T\int _0^t q'(t)q'(s)sin ~\omega (T-t)
2828: sin ~\omega s ~ds ~dt \}
2829: \er
2830: 
2831: As shown in \cite{feyn} the influence functional for this class
2832: of problems has the general form
2833: \be
2834: F(q,q^{\prime})=\int _0^T \int _0^t [q(t)-q'(t)][a(t,s)q(s) - a ^*(t,s)q'(s)]
2835: ~dt~ds
2836: \ee
2837: Thus we can extract $a(t,s)$ from the $q(t)q(s)$ terms.
2838: Calculationally it is easier to look at the $q(t)q^{\prime}(s)$
2839: terms to get $a(t,s)$ as the terms in $S_{cl}$ that are quadratic in $q(t)$ do not contribute. We find for the $\alpha$
2840: independent part of the $q(t)q(s)$ term the following sum of
2841: four terms(an overall factor of ${C^2\over M\omega}$ has
2842: been suppressed for the moment):
2843: 
2844: \[
2845: -{1\over2 sin ^2 ~\omega T} \int _0^T \int _0^t q(t)q(s)sin ~\omega t
2846: sin ~\omega s
2847: \]
2848: \[
2849: -{1\over2 sin ^2 ~\omega T} \int _0^T \int _0^t q(t)q(s)sin ~\omega (T-t)
2850: sin ~\omega (T-s)
2851: \]
2852: \[
2853: -{e^{\i\omega T}\over 2sin^2 ~\omega T}\int _0^T \int _0^t q(t)q(s)
2854: [sin~\omega t sin ~\omega (T-s) +sin~\omega s sin ~\omega (T-t) ]
2855: \]
2856: \be
2857: -{i\over sin ~\omega T}\int _0^T \int _0^t q(t)q(s)sin~\omega (T-t)
2858: sin ~\omega s
2859: \ee
2860: 
2861: These can be rewritten as functions of $t-s$ and $t+s$.
2862: The $t-s$ terms are(after restoring all factors)
2863: \br
2864: &-&{C^2\over M\omega}{1\over 4 sin ^2 ~\omega T}[2cos ~\omega (t-s)
2865: -{e^{-i\omega T}} cos ~\omega (T-s+t)\nonumber\\
2866: &-&{e^{-i\omega T}} cos ~\omega (T-t+s)
2867: -2i\sin{\omega T}cos ~\omega (T-t+s)]
2868: \er
2869: They add up to
2870: \be \label{GS}
2871: -{C^2\over M\omega}[{cos ~\omega (t-s)\over 2} + i {sin ~\omega (t-s)\over 2}]
2872: = -{C^2\over 2M\omega}e^{i\omega (t-s)}
2873: \ee
2874: This is the influence functional when the $Q$-oscillator is in
2875: its ground state \cite{feyn}.
2876: We also have to check that the $t+s$ terms
2877: are zero
2878: The $t+s$ terms are
2879: \br
2880: &~&{1\over 4 sin ^2~ \omega T}cos ~\omega (t+s) +
2881: {1\over 4 sin ^2~ \omega T}cos ~\omega (2T-t-s)\nonumber\\
2882: &~&{ie^{-i\omega T}\over 2 sin^2 ~ \omega T }cos~ \omega (T-s-t)-
2883: {i\over 2sin \omega T}cos \omega (T-t-s)
2884: \er
2885: 
2886: They can be seen to add up to zero.
2887: 
2888: Now we turn to the $\xi$-dependent terms.First let
2889: us look at the terms quadratic in $\xi$:
2890: \[
2891: {\hbar\over 4M\omega}[({2M\omega\over\hbar})(\xi+\xi^*)^2]={1\over 2}(\xi+\xi^*)^2
2892: \]
2893: the exponential of which exactly cancels the factor ${\cal N}$.
2894: 
2895: \be \label{ES}
2896: \ee
2897: 
2898: Now we look at the terms linear in $\xi$:
2899: \[
2900: {\hbar\over 4M\omega}[2F\sqrt{{2M\omega\over\hbar}}
2901: (\xi e^{-i\omega T}+\xi^* e^{i\omega T})+
2902: 2\sqrt{{2M\omega\over\hbar}}(\xi+\xi^*)
2903: {iC\over\hbar \sin{\omega T}}\int_0^T \sin{\omega(T-t)}
2904: (q-q^{\prime})]
2905: \]
2906: On using the expression for $F$ and some rearrangement
2907: it is easy to see that this equals
2908: \be
2909: {iC\over\hbar}\sqrt{{\hbar\over 2M\omega}}\int_0^Tdt~
2910: (q(t)-q^{\prime}(t))
2911: (\xi e^{-i\omega T}+\xi^* e^{i\omega T})
2912: \ee
2913: which can be reexpressed as
2914: \be
2915: iC\xi(\eta -\eta^{\prime})+iC\xi^*(\eta^*-{\eta^{\prime}}^*)
2916: \ee
2917: where
2918: \be \label{eta}
2919: \eta = \sqrt{{1\over 2M\omega\hbar}}\int_0^Tdt~q(t)e^{-i\omega t}
2920: \ee
2921: We can thus express the influence functional for the
2922: case when the $Q$-oscillator is in the coherent state
2923: as
2924: \be
2925: F_{\xi}(q,q^{\prime})=F_0(q,q^{\prime})~e^
2926: {iC\xi(\eta -\eta^{\prime})+iC\xi^*(\eta^*-{\eta^{\prime}}^*)}
2927: \ee
2928: with $\eta$ given by eqn (\ref{eta}).
2929: Within perturbation
2930: theory we can expand the exponential of the linear
2931: functional in $q,q^{\prime}$.It turns out that the
2932: linear term does not contribute and the quadratic
2933: terms are:
2934: $$
2935: -{1\over 4M^2\omega}[\xi^2~(\eta-\eta^{\prime})^2
2936: +{\xi^*}^2~(\eta^*-{\eta^{\prime}}^8)^2
2937: +2|\xi|^2~|(\eta-\eta^{\prime})|^2]
2938: $$
2939: It is sufficient to look at terms of the type
2940: $$
2941: \int_0^T~dt\int_0^t~ds {\Delta a(t,s)}_{coh} q(t) q(s)
2942: $$
2943: to determine the {\em one complex function}
2944: ${\Delta a(t,s)}_{coh}$
2945: that determines the additional effective influence functional.
2946: After some algebra it follows that
2947: \be
2948: {\Delta a(t,s)}_{coh}~=~{C^2\over 4M^2\omega}~
2949: [2\xi^2 e^{-i\omega(t+s)}+2{\xi^*}^2 e^{i\omega(t+s)}+
2950: 2|\xi|^2 (e^{i\omega (t-s)}+e^{-i\omega(t-s)}]
2951: \ee
2952: Now we turn to the evaluation of the effective influence
2953: functional when the $Q$-oscillators are in the $m$-th
2954: number state $|m\ran $. For this we recall the
2955: expansion of the coherent state $|\xi\ran$ in terms
2956: of the number states:
2957: \be
2958: |\xi\ran=e^{-|\xi|^2/2}\sum_n~{\xi^n\over \sqrt{n!}}|n\ran
2959: \ee
2960: Thus
2961: \be
2962: F_{\xi}(q,q^{\prime})=e^{-|\xi|^2}\sum_{n,m}
2963: {\xi^n\over\sqrt{n!}}{{\xi^*}^m\over \sqrt{m!}}~
2964: F_{m,n}(q,q^{\prime})
2965: \ee
2966: Therefore to get the influence functional(in the
2967: perturbative sense described above) for the case in
2968: which the $Q$-oscillators are in the state $|m\ran$
2969: one has to expand $e^{|\xi|^2}~F_{\xi}$ in
2970: powers  of $|\xi|^2$(and quadratic in $q(t)$
2971: and pick the coefficient of
2972: $|\xi|^{2m}$.The $m$-th term of the expansion is
2973: \be
2974: {|\xi|^{2m}\over m!}~a_0(t,s)
2975: +
2976: {|\xi|^{2(m-1)}\over (m-1)!}~ C^2 |\eta|^2
2977: \ee
2978: It is elementary to show that
2979: \be \label{C28}
2980: a_m(t,s)=a_0(t,s)+m{C^2\over M\omega}(e^{i\omega(t-s)}
2981: +e^{-i\omega(t-s)})
2982: \ee
2983: 
2984: \begin{thebibliography}{999}
2985: 
2986: 
2987: \bibitem{KRBS} S. Kalyana Rama and B. Sathiapalan,
2988: {\em On the Role of Chaos in the AdS/CFT Connection}
2989: Int. J. Mod. Phys. A14 (1999) 2635, hep-th/9905219.
2990: 
2991: \bibitem{maldacena} J. Maldacena, Adv. Theor. Math. Phys.
2992: {\bf 2}(1998), hep-th/9711200.
2993: 
2994: \bibitem{witten} E. Witten, Adv. Theor. Math. Phys.
2995: {\bf 2}(1998), hep-th/9802150.
2996: 
2997: \bibitem{reichl} L.E. Reichl, "A Modern Course in 
2998: Statistical Physics", University of Texas Press, 
2999: Austin, p.241.
3000: 
3001: \bibitem{Sinai} Ya. G. Sinai in the {\it The Boltzmann
3002: Equation},ed. E.G.D. Cohen and W. Thirring(Springer-
3003: Verlag, Vienna, 1973).
3004: 
3005: \bibitem{ehren} Paul and Tatiana Ehrenfest,{\it The
3006: Conceptual Foundations of the Statistical Approach
3007: in Mechanics},Cornell University Press,Ithaca,New
3008: York,1959.
3009: 
3010: \bibitem{gibbs}J.W. Gibbs,"Elementary Principles in
3011: Statistical Mechanics", Dover Edition, 1960.
3012: \bibitem{arnold}V.I. Arnold, "Mathematical Foundations
3013: of Classical Mechanics", Springer-Verlag,1980, p.71.
3014: 
3015: \bibitem{chandra}S. Chandrasekhar, Rev. Mod. Phys. {\bf 15} 1(1943).
3016: 
3017: \bibitem{JvN}J.v. Neumann, Zeitschrift fur Physik,
3018: {\bf 57}, 30, 1929.
3019: 
3020: \bibitem{pauli}Sommerfeld Festschrift(Leipzig 1928)
3021: p.30; W. Pauli and M. Fierz, Zeits. Phys.{\bf 106}
3022: 572{1937}.
3023: 
3024: \bibitem{kampen}N.G. Van Kampen {\it Fundamental 
3025: Problems in Statistical Mechanics of Irreversible
3026: Processes} in {\it Fundamental Problems in Statistical
3027: Mechanics}, ed.E.G.D. Cohen,North-Holland Publishing Company,1962.
3028: 
3029: \bibitem{Biro}T. S. Biro, S. G. Matinyan and B. Mueller,
3030: ``Chaos and Gauge Field Theory'', (1994),World Scientific, Singapore ;
3031:  T. S. Biro, C. Gong and B. Mueller, Phys. Rev. D52 (1995) 1260,
3032: hep-ph/940932.
3033: 
3034: \bibitem{Muel} B. Mueller and A. Trayanov, Phys. Rev. Lett. 68 (1992) 3387 ;
3035: T.S.Biro, M. Feurstein, H. Markum, hep-lat/9711002.
3036: 
3037: \bibitem{F1}
3038: V.V.Flambaum and F.M.Izrailev,
3039: {\em Time dependence of occupation numbers and thermalization
3040: time in closed chaotic many-body systems},
3041: quant-ph/0108109.
3042: 
3043: \bibitem{FI2}
3044: V.V.Flambaum and F.M.Izrailev,
3045: {\em Entropy production and wave packet dynamics in
3046: the Fock space of closed chaotic many-body systems},
3047: quant-ph/0103129.
3048: 
3049: \bibitem{I}
3050: F.M.Izrailev,
3051: {\em Quantum Chaos and Thermalization for Interacting Particles},
3052: lectures given in the CXLIII Course "New Directions
3053: in Quantum Chaos" on the International School of
3054: Physics "Enrico Fermi"; Varenna, Italy, July 1999;
3055: to be published in Proceedings;
3056: cond-mat/9911297.
3057: 
3058: \bibitem{CIK}
3059: Doron Cohen, Felix M. Izrailev, and Tsampikos Kottos,
3060: {\em Wavepacket dynamics in energy space, RMT and
3061: quantum-classical correspondence},
3062: Phys. Rev. Lett. {\bf 84} (2000) 2052;
3063: chao-dyn/9909015.
3064: 
3065: \bibitem{FGGP}
3066: V. V. Flambaum, A. A. Gribakina, G. F. Gribakin, and
3067: I. V. Ponomarev,
3068: {\em Interaction-Driven Equilibrium and Statistical Laws
3069: in Small Systems. The Cerium Atom},
3070: Phys. Rev. {\bf E 57} (1998) 4933;
3071: cond-mat/9711213.
3072: 
3073: \bibitem{BGIC}
3074: F.Borgonovi, I.Guarneri, F.M.Izrailev, and G.Casati,
3075: {\em Chaos and Thermalization in a Dynamical Model of
3076: Two Interacting Particles},
3077: chao-dyn/9711005.
3078: 
3079: \bibitem{FI3}
3080: V.V. Flambaum and F.M. Izrailev,
3081: {\em Distribution of occupation numbers in finite
3082: Fermi-systems and role of interaction in chaos
3083: and thermalization},
3084: Phys. Rev. {\bf E 55} (1997) R13; 
3085: cond-mat/9610178.
3086: 
3087: \bibitem{GOM}
3088: Jochen Gemmer, Alexander Otte, and Guenter Mahler,
3089: {\em Quantum Approach to a Derivation of the Second
3090: Law of Thermodynamics},
3091: quant-ph/0101140.
3092: 
3093: \bibitem{S}
3094: D. L. Shepelyansky,
3095: {\em Quantum Chaos \& Quantum Computers},
3096: Physica Scripta {\bf T90} (2001) 112;
3097: quant-ph/0006073. 
3098: 
3099: \bibitem{GS}
3100: B. Georgeot and D. L. Shepelyansky,
3101: {\em Emergence of Quantum Chaos in Quantum Computer Core
3102: and How to Manage It},
3103: quant-ph/0005015.
3104: 
3105: \bibitem{JS}
3106: Ph. Jacquod and D. L. Shepelyansky,
3107: {\em Emergence of quantum chaos in finite interacting
3108: Fermi systems},
3109: cond-mat/9706040.
3110: 
3111: \bibitem{LMI1}
3112: G. A. Luna-Acosta and J. A. Mendez-Bermudez, F. M. Izrailev,
3113: {\em Periodic Chaotic Billiards: Quantum-Classical
3114: Correspondence in Energy Space},
3115: Phys. Rev. {\bf E 64} (2001) 036206;
3116: cond-mat/0105108. 
3117: 
3118: \bibitem{LMI2}
3119: G. A. Luna-Acosta, J. A. Mendez-Bermudez, and F. M. Izrailev,
3120: {\em Quantum-classical correspondence for local density of
3121: states and eigenfunctions of a chaotic periodic billiard},
3122: Phys. Lett. {\bf A 274} (2000) 192-199;
3123: nlin.CD/0002044.
3124: 
3125: \bibitem{MS}
3126: B. Mehlig and M. Santer,
3127: {\em Universal eigenvector statistics in a quantum
3128: scattering ensemble},
3129: cond-mat/0012025.
3130: 
3131: \bibitem{P}
3132: Don N. Page,
3133: {\em Average Entropy of a Subsystem},
3134: Phys.Rev.Lett. {\bf 71} (1993) 1291;
3135: gr-qc/9305007.
3136: 
3137: \bibitem{SS}
3138: Siddhartha Sen,
3139: {\em Average Entropy of a Subsystem},
3140: Phys.Rev.Lett. {\bf 77} (1996) 1;
3141: hep-th/9601132.
3142: 
3143: \bibitem{sred}
3144: Mark Srednicki,
3145: {\em The approach to thermal equilibrium in quantized
3146: chaotic systems},
3147: J. Phys. {\bf A 32} (1999) 1163, 
3148: cond-mat/9809360;
3149: {\em Thermal Fluctuations in Quantized Chaotic Systems},
3150: J.Phys. {\bf A 29} (1996) L75,
3151: chao-dyn/9511001;
3152: {\em Does Quantum Chaos Explain Quantum Statistical Mechanics?},
3153: cond-mat/9410046;
3154: {\em Quantum Chaos and Statistical Mechanics},
3155: talk given at the Conference on Fundamental Problems
3156: in Quantum Theory, Baltimore,
3157: cond-mat/9406056;
3158: {\em Chaos and Quantum Thermalization},
3159: Phys. Rev. {\bf E 50} (1994) 888,
3160: cond-mat/9403051.
3161: 
3162: \bibitem{Berry}
3163: M. V. Berry,
3164: J. Phys. {\bf A 10} (1977) 2083;
3165: In Les Houches XXXVI,
3166: {\em Chaotic Behaviour of Deterministic Systems},
3167: Edited by G. Iooss, R. H. G. Helleman, and R. Stora,
3168: North-Holland, Amsterdam, 1983;
3169: In Les Houches LII,
3170: {\em Chaos and Quantum Physics},
3171: Edited by M. -J. Giannoni, A. Voros, and J. Zinn-Justin,
3172: North-Holland, Amsterdam, 1991.
3173: 
3174: \bibitem{Sinha}
3175: Supurna Sinha,
3176: {\em Decoherence at Absolute Zero}, 
3177: Phys. Lett. A 228 (1997) 1. 
3178: 
3179: 
3180: \bibitem{G}
3181: M. C. Gutzwiller,
3182: {\em Chaos in Classical and Quantum Mechanics},
3183: Springer-Verlag (1990).
3184: 
3185: \bibitem{HS}
3186: H. J. Stockmann,
3187: {\em Quantum Chaos, an Introduction},
3188: Cambridge University Press (1999).
3189: 
3190: 
3191: 
3192: \bibitem{Haake}
3193: F. Haake,
3194: {\em Quantum Signatures of Chaos},
3195: Springer Series in Synergetics, 54,
3196: Springer-Verlag (2000), Second edition.
3197: 
3198: \bibitem{feyn}
3199: R. P. Feynman and A.R. Hibbs,"Quantum Mechanics and 
3200: Path Integrals",McGraw-Hill Publishers,1965.
3201: 
3202: \end{thebibliography}
3203: 
3204: \end{document}
3205: 
3206: 
3207: 
3208: 
3209: 
3210: 
3211: 
3212: 
3213: \\
3214: Title: On the Emergence of the Microcanonical Description
3215: from a Pure State
3216: Authors: N. D. Hari Dass, S. Kalyana Rama and B. Sathiapalan 
3217: Cooments: 58 pages, 2 figures. A reference added. 
3218: Report-no: IMSc/2001/12/55
3219: Subj-class: Statistical Mechanics 
3220: \\
3221: We study, in general terms, the process by which a pure state can
3222: ``self-thermalize'' and {\em appear} to be described by a
3223: microcanonical density matrix.  This requires a quantum mechanical
3224: version of the Gibbsian coarse graining that conceptually underlies
3225: classical statistical mechanics. We introduce some extra degrees of
3226: freedom that are necessary for this. Interaction between these degrees
3227: and the system can be understood as a process of resonant absorption
3228: and emission of ``soft quanta''. This intuitive picture allows one to
3229: state a criterion for when self thermalization occurs. This paradigm
3230: also provides a method for calculating the thermalization rate using
3231: the usual formalism of atomic physics for calculating decay rates. We
3232: contrast our prescription for coarse graining, which is somewhat
3233: dynamical, with the earlier approaches that are intrinsically
3234: kinematical.  An important motivation for this study is the black hole
3235: information paradox.
3236: \\
3237: 
3238: