• Nenhum resultado encontrado

Quantum ohmic dissipation : coherence vs. incoherence and symmetry-breaking. a simple dynamical approach

N/A
N/A
Protected

Academic year: 2024

Share "Quantum ohmic dissipation : coherence vs. incoherence and symmetry-breaking. a simple dynamical approach"

Copied!
16
0
0

Texto

(1)

HAL Id: jpa-00210151

https://hal.archives-ouvertes.fr/jpa-00210151

Submitted on 1 Jan 1985

HAL

is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire

HAL, est

destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés.

Quantum ohmic dissipation : coherence vs. incoherence and symmetry-breaking. a simple dynamical approach

C. Aslangul, N. Pottier, D. Saint-James

To cite this version:

C. Aslangul, N. Pottier, D. Saint-James. Quantum ohmic dissipation : coherence vs. incoherence and symmetry-breaking. a simple dynamical approach. Journal de Physique, 1985, 46 (12), pp.2031-2045.

�10.1051/jphys:0198500460120203100�. �jpa-00210151�

(2)

Quantum ohmic dissipation : coherence

vs.

incoherence and

symmetry-breaking. A simple dynamical approach

C.

Aslangul (1),

N. Pottier

(1)

and D. Saint-James

(2)

(1)

Groupe de

Physique

des Solides de l’Ecole Normale

Supérieure

(*), Université Paris VII, 2 place Jussieu, 75251 Paris Cedex 05, France

(2)

Laboratoire de

Physique Statistique,

Collège de France, 11 place Marcelin Berthelot,

75231 Paris Cedex 05, France

(Reçu le 7 mai 1985, accepte sous

forme difinitive

le 22 août 1985)

Résumé. 2014 La

dynamique

d’une

particule

située dans un double

puits

de

potentiel

et couplée à un bain de

phonons

suivant une loi de

dissipation

«

ohmique

» est étudiée à l’aide de méthodes usuelles de

mécanique statistique quantique.

Nous complétons les résultats

déjà

connus sur la brisure de

symétrie

à T = 0 de la manière suivante :

(i) en déterminant le comportement

critique

du paramètre d’ordre, on montre que la brisure de

symétrie

n’est complète que pour un

couplage

infini ;

(ii) la

dynamique

suit une loi de Kohlrausch

généralisée

de la forme

exp( 2014 atb),

où les deux paramètres a et b

sont calculés exactement; un ralentissement critique est très clairement mis en évidence.

A température finie, la brisure de

symétrie

ne se

produit

plus, bien

qu’un

comportement précurseur de la tran-

sition se manifeste lorsque T ~ 0+.

Ces résultats sont

généralisés à

une

particule

dans un potentiel à puits multiples : en présence de dissipation ohmique, un

phénomène

de localisation peut se

produire

à T = 0.

En dernier lieu, nous examinons d’autres lois de

dissipation

(non

ohmiques)

et nous montrons que seul le modèle

ohmique

est

capable

de

produire

une telle transition, bien que, là encore, un comportement précurseur existe

pour des lois de

dissipation quasi ohmiques.

Abstract 2014 By

using

standard

quantum-statistical

methods, we

study

the

dynamics

of a

particle

in a double-

well

potential, coupled dissipatively,

in an « ohmic » way, to a bath of phonons.

We refine

previous

results about the T = 0 symmetry

breaking :

(i) the critical behaviour of the order parameter is exhibited and the symmetry is shown to be only

fully

broken

for infinite

coupling;

(ii) the

dynamics

follows a

generalized

Kohlrausch law of the type

exp(- atb),

where both parameters a and b

are exactly calculated; a critical

slowing

down is very

clearly displayed

At finite temperature, the symmetry

breaking

does

disappear,

although a precursor behaviour of the transition exists when T ~ 0+.

These results are extended to a

particle

in a

multiple-well

potential for which a localization

phenomenon

is

seen to occur at T = 0.

Finally,

we investigate other

dissipation

laws (i.e. non ohmic) and we show that the ohmic model is

unique by

its

ability

to

produce

such a transition, although a precursor behaviour can be obtained for near-ohmic

dissipation

laws.

Classification

Physics

Abstracts

03.b5 - 05.30 - 74.50 - 71.50

1. Introduction

The

question

which we intend to discuss in the pre- sent paper is that of the influence of

dissipation

on

(*) Laboratoire associe au C.N.R.S.

the fundamental quantum mechanism of quantum coherence. The

underlying

idea is to decide whether

dissipation destroys phase

coherence.

This

question

is

evidently

related to the

general problem

of relaxation in

quantum systems,

for

which,

as it is well

known,

the lines of attack fall into two

Article published online by EDP Sciences and available at http://dx.doi.org/10.1051/jphys:0198500460120203100

(3)

main

categories [1] :

either

people

have tried new

schemes of

quantization

or

they

have used standard quantum

mechanics, considering

the evolution of the

(small)

system and of the reservoir as a whole.

As underlined in reference

[1],

the first type of

approach

contains either controversial results or

obscurities. In the second type of

approach,

known

as the

system-plus-reservoir approach,

and

pioneered by Senitzky [2],

the interaction of the

system

of interest with the reservoir is

explicitly

taken into account

Irreversible behaviour of the small

system

arises when the reservoir is considered as a continuum of

modes,

thus

rejecting

Poincar6 times to

infinity.

The new aspect of

dissipation

which we would like

to

study

here is the

following :

since

1982,

it has been shown

by

several authors

[3, 4] that,

for a certain

type

of

dissipation (the

so-called « ohmic »

dissipa- tion),

a

spontaneous symmetry breaking

may appear in some

dissipative quantum

systems when the dissi-

pation

becomes

greater

than a critical value.

This is all the more

interesting

as, in these « ohmic »

dissipative quantum

systems, the

Heisenberg

repre- sentation of a certain

coordinate q

follows a classical

equation

of motion

which is

generally

believed to be

obeyed by

macros-

copic

quantum variables

[5] (M

is the mass of the

particle

submitted to a

potential Y(q) ;

q is a

pheno- menological damping

constant or friction coefficient and

F(t)

is a

fluctuating force).

These

systems

have

been the centre of recent interest in the context of

testing

the

applicability

of quantum mechanics to

macroscopic

variables

[6].

Following

references

[3, 4],

we shall

mainly study

here a

particle

in a double-well

potential, coupled dissipatively (in

an « ohmic »

way)

to its environment This

problem

can be very

clearly

settled

down,

but

it cannot be

given

an exact solution. Thus

approximate

treatments have been

proposed

in the literature.

Most of them use either

path-integral Feynman

for-

mulation in

imaginary

time and instanton

techniques,

or renormalization group

approachs [3, 4, 7].

Two

other

approaches

must

equally

be

quoted :

that of

reference

[8],

in which a variational method of dis-

playing

the transition is

proposed,

and that of refe-

rence

[9]

in which a linear response calculation is

handled

As for the real-time

dynamics,

it has not

been

investigated

in

detail, except

for references

[9, 10]

outside the critical

region.

Our purpose in this paper is thus to refine and to

supplement previous

treatments of this

problem,

and

that will be attained

by using

standard quantum- mechanical

theory. Indeed, by using

standard

methods,

we recover

qualitatively

and

quantitatively nearly

all

previous results,

and in

particular

the spontaneous

symmetry breaking

which occurs at zero temperature

[11].

The

improvements

over

previous

treatments

are the

following :

we are able to

give

a

description

of the real-time

dynamics

of the

particle,

for zero

as well as for finite temperature; in the

vicinity

of the

transition

point,

we find a critical

slowing

down.

Another

advantage

of our methods is that the

approxi-

mations involved are more

transparent

from a

physi-

sical

point

of view. Moreover it will be shown that

the order parameter is continuous and reaches its maximum value

only

for infinite

coupling;

the par- ticle is then

fully

localized.

The paper is

organized

as follows. In section

2,

we describe the model under

study;

in section

3,

we

present

an

approximate method,

based on a

master

equation formalism,

which

yields

the real-

time

dynamics

in the case of intermediate or strong

coupling.

In section

4,

we present another

approxi-

mation

scheme,

based on a one-boson

assumption,

which

yields

the real-time

dynamics

in the weak cou-

pling

case. Section 5 contains a

comparison

between

our results and

previous

results on the real-time

dyna-

mics

[9, 10]. Then,

in section

6,

we discuss the

speci- ficity

of the ohmic

dissipation

model. In section

7,

we show how the results on the

particle

in a double-

well

potential

can be extended to a

particle

in a

multiple

well.

Finally,

in section

8,

we present our conclusions.

2. The model

2 .1 THE DAMPED PARTICLE IN THE SYMMETRIC DOUBLE- WELL. - As indicated in the

Introduction, following

references

[3, 4],

we consider a

particle

of mass M

moving

in a

symmetric

double-well

potential V(q) = V( - q)

with its minima located at q =

± q°

2 and

separated by

a local maximum at q = 0.

This

particle

is in contact with a

dissipative medium,

which is

generally

modelled

by

a

phonon

reservoir

of bandwidth

liwe

,

[5].

As

previously indicated,

we

shall

mostly

be interested here in the case of« ohmic »

dissipation,

in which the coordinate of the

particle

evolves with time

according

to the classical

equation

of motion

(1).

The

problem

is the

following :

at time t =

t°,

a

particle

is located in one of the two

wells;

the

tempe-

rature is

supposed

to be low

enough

so that thermal

activation over the

potential

barrier can be

neglected

The

question

arises : how does its average

position

evolve with time ?

Evidently,

a

purely

classical par-

ticles, initially

located in one of the two

wells,

remains

localized on the same side for an infinite time. On the

contrary,

a

purely quantum-mechanical particle (wi-

thout

bath), initially

located in one of the two

wells,

can tunnel

through

the

potential barrier,

and thus

oscillates back and forth between the two wells. This is the

phenomenon

of «

quantum

coherence ». Such

a

particle

can be said to be delocalized : its average

position

is

equal

to zero. Let us now consider a quan-
(4)

tum-mechanical

particle coupled

to a

bath,

in such a

way so that its

position

operator in the

Heisenberg picture obeys

the

equation

of motion

(1).

In this situa-

tion, Chakravarty [3]

and

Bray

and Moore

[4]

have

found a transition : for a weak

coupling

with the

bath,

the

regime

is akin to the pure

quantum regime

and the

particle undergoes

a

damped

oscillation between the two

wells;

for

higher

values of the cou-

pling,

the mean coordinate of the

particle

relaxes

towards zero without

oscillating,

but when the

coupl- ing

to the bath becomes

greater

than a critical

value,

a

spontaneous

symmetry

breaking

occurs, which here

signifies

that the mean coordinate of the

particle eventually

relaxes towards a non-zero

value,

indicat-

ing

a localization in one of the two wells. This

regime

thus bears a resemblance to the classical

regime

and

corresponds

to a

complete

loss of

phase

coherence

[12, 13].

In the

system-plus-reservoir approach,

one

begins by writing

down the total Hamiltonian for this pro- blem.

Following

Caldeira and

Leggett [5],

the dissi-

pative

interaction of the

system

with its environment is modelled

by

a linear

coupling

with a bath of har- monic oscillators. In obvious notations :

The last term has been added in order to cancel

spurious unphysical divergences;

it ensures that there

is not shift in the

potential

due to the

coupling [5].

It is easy to derive from

equation (2)

the

equations

of motion for the coordinate

q(t)

and the momen-

tum

p(t)

of the

particle

in the

Heisenberg picture.

As we

already

shown in a

previous

paper about the

damped

free

particle [14],

once

p(t)

is

eliminated,

the

dynamical equation

for

q(t)

reads as follows :

where the interaction with the bath manifests

itself,

as it is

well-known, by

a

dissipative

retarded term and

by

a

fluctuating

force term, both

being

related

by

the

fluctuation-dissipation

theorem. Note that

equation (3),

which has been established without any

approxi- mation,

contains the exact

dynamics

of the

particle.

Let us now discuss the conditions under which it can

be

given

the form of the classical

equation

of motion

(1).

2.2 THE OHMIC DISSIPATION MODEL. - As

usual,

the

bath is described in the continuous limit

by

its

density

of modes

p(c~). Following

the same line of

reasoning

as in reference

[14]

in the case of the

damped

free

particle,

a classical

equation

of motion of the

type

JOURNAL DE PHYSIQUE. - T. 46, No 12, DtcaoRE 1985

of

equation (1)

for

q(t)

is obtained

by setting

down

where

f,((olco,,)

is some cut-off function such that

£(0)

=

1,

and

decreasing

on a

frequency

range of order co,,. When the choice

(4)

is

made,

it can be shown

[14]

that

K(t)

is a memory kernel

essentially given by

the Fourier transform of the cut-off function

fc(ro/wc)

and

that,

for temperatures such that

T >

liwc, F(t)

is a

rapidly fluctuating

force of

correlation time of order -

coc

1. The constraint

(4) imposed

on the bath

density

of modes and on the

coupling

constant in order to obtain a

classical-like equation

of motion will from now on be referred to as the ohmic

dissipation

model. Let us

emphasize that,

in accordance with the

fluctuation-dissipation theorem,

not

only

the

friction,

but also the

fluctuating

force term have a determined form.

2.3 REDUCTION TO A SPIN-BOSON HAMILTONIAN

[4]. -

Let us consider the

particle

in the double-well and let

S~o

denote the

frequency

of small oscillations around one of the

minima,

and

Yo

the barrier

height

We shall limit the

study

to the

low-temperature

situation

T

V~,

in which thermal activation

over the

potential

barrier can be

neglected Moreover,

we shall suppose that this barrier is

sufficiently high

and

large (Yo > hlmq’)

for the

splitting

coo between the

symmetric

and

antisymmetric

states

(tunnelling frequency)

to be much lower than

00. Finally,

we

shall suppose that the

temperature

is such that

kB T ~~o :

the system can then be

safely

truncated

to

only

one level in each well.

After this

truncation,

we are left with an

effectively

two-level

system coupled

to a bath of

phonons;

this ensemble can be described

by

the

spin-boson

Hamiltonian

(6x,

ay

and a,,

are the standard Pauli

matrices) :

where the

operator §

qo Uz

corresponds

to the

posi-

tion operator q of the

particle ;

in order to be

complete,

we have now to

identify

the

coupling

term of the

spin-

boson Hamiltonian

Hsb (5)

with the

coupling

term

of the Hamiltonian

(2).

This can be done

by identifying

The ohmic constraint

(4)

then takes the form
(5)

where the dimensionless

parameter

a is related to the friction

coefficient by

The cut-off

frequency

We has to be chosen such that

c~~ >

wo and

cor >> kB Tlh. Clearly,

with these assump-

tions,

one

expects

that the

precise

form

of fr

has no

bearing

on the

dynamics

of the

system

in the time domain

c~~ t >

1 and may be chosen at convenience.

To conclude this

section,

let us now comment

briefly

on the reduction of the Hamiltonian

(2)

to

the

spin-boson

Hamiltonian

(5). Clearly,

the operator

2’ qo u%

possesses

only

two discrete

eigenvalues

±

2 Z - 2

whereas the

genuine position operator q

has a conti-

nuous

spectrum.

The kinetic and

potential

energy have been contracted into the u x term, and

since,

as

well-known,

the Pauli matrices

algebra

is

comple- tely

different from the

algebra

of the

position

and

momentum

operators (in particular

it is

impossible

to find an

operator having

the same commutation relation

with a.

than p with

q),

this amounts to carry out a kind of

coarse-graining

in which some infor-

mation has been

lost;

one may

loosely speak

of a

reduction of the Hamiltonian.

Conversely,

the results obtained on the

dynamics

of a. can be cast in terms

of q,

but one cannot

expect u%(t)

to

obey

any classical- like

equation

of motion of

type (1), which,

as

explained above,

is obtained for a

quantum system coupled

to a

bath when

writing

down the

equations

of motion

for both

q(t)

and

p(t).

In the

following

two

sections,

we shall

successively analyse

the cases of weak and moderate or strong

coupling.

3. Intermediate or

strong coupling

case.

3.1 PARTIAL DIAGONALIZATION OF THE HAMILTONIAN.

- As

pointed

out

by Silbey

and Harris

[8]

and

by Zwerger [9],

the

unitary

transformation

diagonalizes

the Hamiltonian

Hgb(5)

when Wo = 0.

The transformed

Hamiltonian n

=

SHsb S -1

reads :

The

operators Bt

are defined as

Formally, J7

divides into an interaction term

and a bath term.

In what

follows,

we shall be interested in the time evolution of the average value of 6z, defined in the

Schrodinger picture

as

(p(t)

is the full

(i.e. system plus bath) density

matrix

and the trace operator acts in the whole

space).

The basic remark is the

following :

the

unitary

transformation S leaves u %

invariant, and,

conse-

quently,

one

gets

with

Since a,,

is a pure

spin

operator,

equation (14)

takes

the form

where the trace

operator

acts in the

spin

space and

p,(t)

is the reduced

spin density

matrix

(the

trace

operator

acts in the bath

space).

Similarly,

the time

derivative åz(t) >

is

given by

By setting

one

gets

The time evolution

of p

has thus been calculated

by using 17

instead of

H,

this

being legitimate

as

long

as one is interested in

quantities

such as a.

which are invariant under the

unitary

trans-

formation S.

3.2 MASTER EQUATION FOR THE REDUCED SPIN DENSITY MATRIX

ps(t).

-

According

to the

general damping

theory [15],

an evolution

equation

for the reduced
(6)

spin density

matrix

ps(t)

can be deduced from the

Liouville

equation

for

p(t) by using

for instance standard

projection operators techniques.

One assu-

mes, as

usual,

that the initial

density

matrix

p(to

=

0)

is

factorized,

i.e. that

where pb

is the

equilibrium

bath

density

matrix.

The time rate of

change « Op,10t >>

then

comprises only

two terms : one is a proper evolution term, and the other one is a relaxation term, due to the interaction with the bath. As

pointed

out

by

several

authors

[16-19],

this relaxation term can be

given

two

forms,

either the usual time-convolution or

retarded

form,

or a « time-convolutionless »

(i.e.

non

retarded)

one.

Clearly

both forms are exact as

long

as no

supplementary approximations

are

made;

however,

in what

follows,

we would like to consider the interaction between the

system

and the bath in

H (the Hint term)

as a

perturbation.

The calculation will be restricted to the Born

approximation (1).

It has been

argued [16-19] that,

from the

point

of

view of a

systematic expansion,

the time-convolution- less formalism

prevails

over the usual convolution one, since in the latter

formalism,

a certain amount of rearrangement of the terms is necessary to obtain a

coherent

expansion

to a

given

order. This is the fundamental reason

why,

besides its

greater practical simplicity,

we

preferred

the time-convolutionless for- malism.

Following

van

Kampen [16]

and Hashitsume et al.

[18],

as

long

as two time scales can

clearly

be

delineated,

in other

words,

as

long

as it exists a short time

scaler,,

characteristic of the relaxation

kernel,

and a

long

time scale T., characteristic of the operator of interest

(here

the

density

matrix

Ps(t)),

one can

write,

for

times t > T,, a time-convolutionless evolution equa- tion for

p,(t)

which takes the form of a

systematic expansion

with

respect

to the

perturbation

where (’" ~b

stands for

Trb

p~’" The time-convolu- tionless

equation

for

ps(t)

takes the

form,

in the Born

approximation

where

Hi’n,(- t’)

is written in the interaction

representation.

It is easy to

give equation (22)

a more

explicit

form

by detailing Hinc

and

H:nt( - t’).

One obtains

where the relaxation functions

0 + + (t), ql - + (t), 0 + - (t)

and

0 - - (t)

are defined as

Remember that the

operators Bt(t)

are

computed

in

the

non-interacting

scheme for the bath.

Equations (22)

to

(24)

constitute

general

relaxation

equations;

(1)

This evidently

implies

that the

perturbation

series properly exists, an assumption which, as

pointed

out

by

V. Hakim, could be dubious in the close

vicinity

of a = 1

(private

communication).

note that

they

are valid whatever the

precise

form of

the

dissipation,

ohmic or non ohmic. Let us now

examine the consequences of the ohmic constraint

(7)

on

the

relaxation

equation

of the reduced

spin

den-

sity

matrix

Ps(t).

3.3 TIME EVOLUTION OF THE REDUCED SPIN DENSITY MATRIX

ps(t)

IN THE OHMIC MODEL. - At

first,

let
(7)

us calculate the average value

Since b] and b,,

are boson operators, one gets

When the ohmic constraint

(7)

is

imposed,

the

exponent in

equation (26)

is seen to

diverge,

for

T = 0,

as well as for finite temperature, so that

( Bt )b

is

equal

to zero in this model. Let us however

emphasize

here that the ohmic model is not the

only

one in

which

( Bt ~b

= 0.

Any

model of

dissipation

in

which

p(w) ~ 1 G(w) 12 ’" c~8

with

6

1 at zero tem-

perature

or

6

2 at finite

temperature

will share this

property.

This

point, and,

more

generally,

the

question

of the

specificity

of the ohmic

dissipation model,

will be discussed later on in detail

(see

Sect.

6).

Thus,

in the ohmic

model,

the relaxation

equation (23)

of the reduced

spin density

matrix

ps(t) considerably simplifies.

It is then

straightforward

to deduce from

equation (19)

the

equation

of motion of

( Qz(t) ~.

This will be achieved in the

following paragraph.

3.4 SPIN DYNAMICS. - In order to obtain a detailed

expression,

it is necessary to calculate the relaxation function

§’s, which,

in the ohmic

model,

are

just time-integrals

of the correlation functions of the operators

B=f:(t), expressed

in the

non-interacting

scheme.

By using

the commutation relations between the Pauli matrices and the

symmetry properties

of

the

§’s,

one gets a closed

equation for ( uz(t) > :

where

A(t)

is defined as

with

and

Equations (27)

to

(30)

constitute our basic equa-

tions ; they

contain

explicitly

all the

dynamics

of the

spin.

Note that the

exponential

time behaviour found

by Chakravarty

and

Leggett ([l0a])

is

simply

obtained

from

equations (27)

and

(28) by rejecting t

to

infinity

in the

integral (28);

in our context, this

corresponds

to a short memory

approximation. Using

the expo- nential cut-off introduced

by

the same authors :

in the ohmic constraint

(7),

one can calculate

simply

the functions

A1(t)

and

A2(t) :

For kB T liwc, expression (32b)

for

A2(t)

reduces to

1 is the « temperature time », as defined

by

Equation (28)

then

yields

The

explicit dynamics of U z( t)

can now be

found

by integrating equation (27).

Let us present and

physically

discuss the results we obtain for

( uz(t) ).

This will be done

successively

for zero and

for finite temperature.

3.4.1

Zero-temperature

case. - When the

tempe-

rature is

equal

to zero,

expression (35)

for

~1(t)

reduces

to

and the solution of

equation (27)

is

easily

obtained

in a closed form :

(Note

that for the

particular

values a =

1/2

and

a = 1 the

preceeding expression

of

Z(t)

is still valid

provided

the limits are

correctly taken.)

In the

meaningful

time domain

c~~ t > 1, L(t)

takes
(8)

the

simpler

form :

This

expression

shows that :

(i)

As

long

as a

1, Z( + oo)

is

equal

to zero. In

terms of the

particle evolving

in the double-well

potential,

this means that its average

position

at

infinite time is

equal

to zero.

Therefore,

in this

regime,

the

particle

can be said to be

delocalized,

albeit no

oscillation does appear.

However,

since for

(X = 0,

the situation

evidently

reduces to the pure quantum situation in which the

particle

oscillates back and forth between the two

wells,

it is

likely

that the present calculation is no more valid for too low values of the

coupling;

this will be discussed later.

(ii)

When a becomes greater than

1, Z( + oo)

takes

a finite

value, given by

The average

position

at infinite time of the

particle evolving

in the double-well

potential

is then different from zero. The

particle

is thus

partly

localized on

the side where it was at time t = 0.

The

degree

of

localization,

which can be charac-

terized

by

the value of

E( + oo),

is a continuous

function of a, which grows very

rapidly (i.e.

on a

region

of width -

(c~o/c~~)2)

towards 1 as soon as

a exceeds 1.

Therefore,

this

system

does

display

a

continuous

symmetry breaking

at the value a = 1 of the

coupling; however,

the symmetry is

fully

broken

(i.e. Z( + oo)

=

1) only

for infinite

coupling,

which seems a

physically

sensible result

(see Fig. 1).

Fig. 1. - Variation of E( + oo) as a function of the inverse coupling constant

03B1-1(w0/wc

= 0. 1). The inset is an enlarg-

ment of the critical region.

The

symmetry breaking

can be described in the

phase

transition

language,

the inverse

coupling

para- meter

a-1 playing

the role of the temperature, and

E( + cc) acting

as the order parameter. This

phase

transition is of infinite order

(see Eq. (39)),

and the

width of the critical

region

is of order

(wOlwe)2,

i.e.

very narrow. This is to be

compared

with the result of references

[3, 4]

in which a discontinuous transi- tion at a = 1 is found. Note however

that,

for we -. oo,

E( + oo)

in our calculation will also present a

jump

at a = 1.

The

time-dependence of E(t)

reveals very

interesting

features. The full

time-dependence

of

Z(t)

is

given by equation (37); however,

it is sufficient to

analyse

formula

(38)

which

corresponds

to the

meaningful

domain

We t

> 1. The

following

characteristics appear :

(i)

As

long

as a

1/2, E(t)

relaxes towards zero more

rapidly

than a pure

exponential,

i.e.

Z(t) - exp( - atb),

b > 1.

(ii)

For a =

1/2, Z(t)

has a

purely exponential

behaviour.

(iii)

When

1/2

a

1, ~(t)

has a stretched expo- nential form : in other

words,

it evolves

according

to the so-called Kohlrausch law

Z(t) - exp( - atb),

0 b 1. Such laws are believed to occur in a wide range of

phenomena

and

materials,

and to contain

the essential

physics

of relaxation in

complex, strongly interacting

materials

[20, 21].

This law appears here

as a consequence of the ohmic

dissipation model,

and manifests itself in a certain range of values of the

coupling

constant.

(iv)

As a

approaches

1 from

below,

the relaxation of

Z(t)

towards zero is slower and slower. At

a=1,

the relaxation is seen to take the

powerlaw

form

(1 + (o2 t2)- C02/2(02 ;

since

~o/~c ~ 1,

it is

extremely slow;

this can be viewed as a critical

slowing

down.

(v)

For a >

1, ~(t)

starts from 1 and goes uni-

formly

very

slowly

to its final

value % 1 ; Z(t)

behaves

as

exp(- atb),

b 0.

So in all cases, one can say that the relaxation of

Z(t)

follows for

we t

> 1 a «

generalized >>

Kohlrausch

law;

the

corresponding

curves are

plotted

on

figure

2.

To

give

an idea of the

extremely

slow character of the evolution of

Z(t)

when a >

1,

let us indicate

some orders of

magnitude.

For

instance,

when

a =

1.02, E( + oo)

= 0.786 and

E(we t

=

1010) =

0.865. If c~~ ~ 1013

S- 1,

which is a

typical phonon

(9)

Fig. 2. - Variation of E(t) at T = 0 as a function of t for various values of

a(mo/m~

= 0.1).

Actually,

for this value of

wo/w~,

the curve a = 0.1 is

slightly

outside the region

of

validity

and is only

given

here for the sake of illustration.

frequency,

this

corresponds

to a

macroscopic

time

t ~

10- 3

s, for which

L(t)

is still

relatively

far from

its final value. This behaviour is

clearly

apparent on

figure 3,

which

displays

the variations of

E(t)

for

a = 1.02 and a = 0.9.

3.4.2 Finite temperature case. - In this case,

A(t)

is

given by equation (35)

and then it is not easy to derive from

equation (27)

a closed form

expression

for

~(t).

Thus one must in

general

resort to a nume-

rical

integration.

However,

an

asymptotic formula,

valid for times

t >

i/2

a, can be obtained.

Indeed, provided

that

is

finite,

the upper bound of the

integral

in

equation (35)

can be extended to

infinity

when t >

1/2

a. One thus

gets in this time domain

or

The rate

A( (0)

can

easily

be calculated :

where r denotes the standard Euler’s gamma func- tion.

Besides,

one can

verify, by directly expanding

the

integral expression for ( uz(t»

which can be

deduced from

equation (35),

that the constant in

equation (41)

is indeed

equal

to

( 7~(0)).

This cal-

culation thus shows

that,

at finite

temperature,

uz(t)

relaxes towards zero in an

exponential

manner as soon as t >

z/2

a.

Moreover, equation (35)

also proves that the effect of

temperature

on the time evolution of

( ~(0 )

is

negligible

as

long

as t

r/2

cx; in other

words,

on

this time range, the behaviour of

( 6z(t) ~

is the same

as at zero temperature.

Fig. 3. - Variation of f(t) at T = 0 as a function of t for

a = 1.02 and a =

0.9(wo jcv~

= 0.1 ). The asymptotic value

for a = 1.02 is equal to 0.786.

Stricto sensu, the

symmetry breaking disappears

at

finite

temperature :

whatever the value of a

(greater

or lower than

1, provided

our calculation should be

valid,

a condition which

only

excludes low values of a, as it will be seen

later,

and

perhaps

the

vicinity

of

1), E(t)

goes to zero as t -~ oo. The

particle

in the

double-well

potential

is

always

delocalized. However, the « relaxation

time »,

as defined

by

has for limit value as To

0,

either 0

(when

a

1/2),

or oo

(when

a >

1/2).

This is related to the corres-

ponding

zero-temperature behaviour

of ( a_,(t) >

as

mentioned above in 3.4. l. For T >

0,

one could

naively expect

that the

decay

time

of ~ a_,(t) >

should

be of the order Of T. However,

equations (41)

and

(43)

show that this time is renormalized

by

a factor

,

which,

for a >

1,

. -, ,

becomes

extremely large

when T -+ 0. The

spin

thus

displays

a behaviour which is a precursor of the zero-

temperature transition at a = 1. It can be seen

that,

at finite

temperature,

the relaxation of

E(t)

takes

place

on a

microscopic

time even when a > 1.

Let us consider for instance the case a Z 1. When We T =

10,

which for We ~

1013 s-1 corresponds

to

a

temperature

T - 2.4

K,

one

gets

we

TR

=

103

or

TR ~ 10-’o

s. Thus it is

only

for

extremely

low

values of the

temperature

that

TR

would take macro-

scopic values;

for

instance,

for T - 2.4 x 10-’

K,

one would get

TR

=

10- 3

s.

The curves

representing uz(t) >

as a function of

t for different values of a are

plotted

on

figure

4.

These curves have been obtained

by

numerical

integration

of

equation (27),

with

~1(t) given by

equa- tion

(35). By comparison

with the

corresponding

T = 0 curves

(Fig. 2),

two trends are observed : for

relatively high

values of the

coupling, uz(t) >

closely

follows the

exponential

law as

given by equation (41),

but for

relatively

small a

values,

the

dynamics

appears to be

quite

similar to the

dynamics

at T = 0. These two trends will be confirmed

by

a
(10)

Fig. 4. - Variation of E(t) at finite temperature

(we

z = 10)

as a function of t for various values of

rx(wolwe

= 0.1).

more detailed

quantitative analysis,

which will be done in the

following paragraph.

3 . S CONDITIONS OF VALIDITY OF THE TREATMENT. -

Let us now

precise

the conditions under which the convolutionless

equation (27)

and the results which

were derived from it in the two

preceding paragraphs

are valid. Our discussion will follow the arguments

developed by

van

Kampen [16]

and

by

Hashitsume

et al.

[18],

and will be carried out

successively

for

zero and for finite temperature.

3.5.1

Zero-temperature

case. - When T =

0,

an

exact

analytical

solution of

equation (27)

can be found

(Eqs. (37)

and

(38)).

Thus it is sufficient to discuss the conditions of

validity

of the convolutionless

equation

itself. As

already indicated,

this

equation

is valid as

long

as two time scales can be

clearly

delineated : the short one, ’rc’ which characterizes the relaxation

kernel,

and the

long

one, zm, which characterizes the operator of interest

(here a,,(t) >).

Since the two functions to be considered are not

just exponential functions,

it is not so

simple

to

assign

to them well-defined time

scales,

all the more

as

they

possess different time

dependences. However,

one can

roughly

define ’rc as the time after which the relaxation kernel

(1 + W2 t2)-, (see Eq. (36))

has

decreased

by

a factor

lie,

which

yields :

One must have i~ im, where im is linked to

z uz(t).

For a >

1,

this condition is

automatically satisfied,

since the time scale

of ~ U z( t) )

is

extremely long,

as indicated above. For a

1, equation (38) approximately yields :

Note that

T- 1

is the so-called renormalized tun-

neling frequency appearing

in

equation (15)

of refe-

rence

[10a].

The

separation

of time scales

implies

that :

and,

for

w~lc~o > 1,

this can be rewritten as :

The

separation

of the two time scales also means

that the renormalized

tunnelling frequency

has to be

small as

compared

to the bare one; in other

words,

the effective

tunneling

time Tm must be great in

comparison

with the relaxation time of the bath as

measured

by

T,.

Our calculation thus has a wide range of

applica- bility since, according

to

equation (47), only

values

of a near zero are excluded. When the condition

(47)

is

satisfied,

the convolutionless

equation (27)

and its

solution are valid for

times t >

T,, as

long

as the

expansion

parameter Wo T, is

small,

i.e. :

Note

that,

for

high

a

values,

the

expansion parameter

reduces to

woll

We

(an effectively

small

quantity

in

the

model), whereas,

for small a

values,

it is

approxi- mately equal

to

(wolwe) el/2a,

this latter

quantity being

indeed small as

long

as the

separation

of time

scales,

as

given by equation (47),

is achieved.

3.5.2 Finite temperature case. - The discussion of the

validity

of the convolutionless

equation (27)

at

finite temperature is more

involved; indeed,

the solu-

tion of

equation (27)

has been obtained

through

a

numerical

integration,

which renders uneasy the discussion of the

separation

of time scales.

However,

as shown

above,

in certain cases,

analytical approxi-

mations of the solution can be

given,

which then

simplifies

the discussion.

Let us first define the short time scale at finite tem-

perature. The relaxation kernel

(see Eq. (35))

is the

product

of two

functions,

one of which

being

the

relaxation kernel at T =

0, (1 + COC2 t 2) - a,

of time

scale,r, (Eq. (45)),

and the other one,

(tlt)/sinh (tlt))2a,

decreasing approximately

on a time scale of the order

T/2

a. One can thus

roughly

define the short time scale at finite temperature as min

(~~, r/2 a).

Since for times t >

r/2 a, ( a_,(t) > approximately

relaxes

according

to an

exponential

law of time

constant

TR (see Eq. (41)),

this

approximation

would

be

meaningful

for

nearly

the whole time domain

provided

that

Condition

(49)

is fulfilled in the

region

above the

solid curve 1 of the

(a, T) plane (see Fig. 5).

In this

region,

the condition of

separation

of time scales is

automatically realized,

which in turn insures the

validity

of the convolutionless

equation

itself.

As

explained

in

paragraph 4,

the

~ehaviour

of

( U z( t) >

in the time range t ;5

i/2

a is the same as

at zero temperature.

Therefore,

as

long

as

z Uz(t) >

is well

represented by

its zero

Referências

Documentos relacionados