• Aucun résultat trouvé

Dynamics in a spin glass model with large order parameter fluctuations

N/A
N/A
Protected

Academic year: 2021

Partager "Dynamics in a spin glass model with large order parameter fluctuations"

Copied!
10
0
0

Texte intégral

(1)

HAL Id: jpa-00246360

https://hal.archives-ouvertes.fr/jpa-00246360

Submitted on 1 Jan 1991

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.

Dynamics in a spin glass model with large order parameter fluctuations

A. Jagannathan

To cite this version:

A. Jagannathan. Dynamics in a spin glass model with large order parameter fluctuations. Journal de

Physique I, EDP Sciences, 1991, 1 (5), pp.659-667. �10.1051/jp1:1991160�. �jpa-00246360�

(2)

Classification

Physics

Abstracts

75.10N 75.40G 75.50L

Dynamics in

a

spin glass model with large order parameter

fluctuations

A.

Jagannathan

Centre de

Physique Th60rique (*),

l3cole

Polytechnique,

91128 Palaiseau Cedex, France

(Received

I October 1990, revised J7 December 1990,

accepted

30

January 1991)

Abs«act. A

dynamical

extension of the mean

spherical

model for

spin glasses

is

presented.

Features of

nonselfaveraging

and

large

fluctuations present in the

original

model are seen to result in the existence of a spectrum of time scales of relaxation,

leading

to

non-exponential spin

correlations.

Spherical

models may

arise, generally speaking,

in situations where the

system

under consideration possesses certain local constraints that are difficult to treat

exactly. Then,

a

possible approximation

consists in

replacing

the set of local constraints

by

a

single, global,

constraint. The name

spherical

arises from the first use of this

approximation

in

treating

the

Ising ferromagnet [I]

when Berlin and Kac solved the model

by replacing

the local

Ising spin

variables

by

continuous variables that satisfied a

single global spherical

normalization. In other contexts, one has the slave boson

technique

in models for

strongly

correlated electron

systems,

where the fixed local

(onsite) particle

number constraint is

replaced by

a

global

number conservation of

particles

over the entire lattice

[2].

More

recently,

among models for

Heisenberg antiferromagnets utilizing

the

Schwinger

boson

representation,

the

spherical approximation

has been used to

approximate

the local boson

occupation

number constraint

[3].

In all these cases, it is assumed that the local variables will not be too much affected

by

the

admittedly

extreme

simplification

made

by

the

spherical approximation,

and that fluctuations remain in some sense small. This

assumption

was proven correct in some of the cases above

by

an

explicit

calculation of fluctuations. The interest of this paper lies in the case where this

assumption

is no

longer

true, and this occurs in the use of the

spherical

model for random systems.

The

present study

considers a

possible dynamics

for a

spin glass

system that is

subject

to the weak version of the

spherical length

constraint. The unusual fluctuation

properties

in the

model,

described

below,

are

expected

to

give

rise to unconventional

dynamical properties

in

spite

of the

simplicity

of the model.

Compared

with results obtained for

statics,

the results for its

dynamical properties

are

necessarily

on less firm

ground,

as

they

are model

dependent,

and

one can

certainly

conceive of a number of

quite

different

dynamical

extensions of the

original

(*)

CNRS

Laboratory

UPR 14.

(3)

660 JOURNAL DE PHYSIQUE I N 5

model. Here we

present

what we

hope

to be a

plausible, physically

relevant but not

rigorously justified, dynamics

for the mean

spherical spin glass

model. The motivation is

finally,

of

course, to

gain

some

understanding

of the time

dependent properties

of random systems, which are

typically nonselfaveraging,

have many

coexisting

time scales of

behavior,

and are still

poorly

understood.

The

spherical

model of the infinite range

spin glass

possesses a

phase

transition

analogous

to that of Bose-Einstein condensation in a gas of

noninteracting

Bose

particles.

The

spherical ferromagnet

in dimensions greater than two does so as

well,

about which more is said later. In the case of the bosons it has been shown that the choice of

ensemble,

canonical

(constraining

the

particle

number to be a fixed

value)

or

grand

canonical

(where

the constraint is

applied

to the average number of

particles)

does not affect the

thermodynamic

averages but

gives

rise instead to very different fluctuation

properties

in the condensed

phase [4].

A similar situation arises in the case of the

spin

model. The

analogue

of

relaxing

the

particle

number constraint

is,

in the case of the

spin model,

the relaxation of the strict

spherical spin length

constraint in favour of the mean

spherical

constraint

(where

the constraint is satisfied

only

on the

average).

The latter case

corresponds

to

taking

the

grand

canonical ensemble for the Bose gas, where the average number of

particles

is

fixed, using

a chemical

potential,

but fluctuations around this number occur. In

general,

one

expects

that these fluctuations are

small, typically finite,

in the

thermodynamic limit,

so that total number of

particles

is a well defined

thermodynamic

observable. However this is not the case in the ideal Bose gas, and its

spin analogue,

the

spherical

model. This

property

of the

spin glass

is not

altogether

a trivial

result,

and there are some subtleties involved as will become evident further

below,

when we discuss the effects of external fields and spontaneous symmetry

breaking

in the model.

In the mean

spherical spin glass

model

it

has been shown that there is

a nontrivial distribution of

possible

values of the order

parameter,

which is the

expectation

value of a certain

appropriately

defined

spin

variable

[5].

In contrast to the

corresponding Ising

or S~

models,

or indeed the strict

spherical model,

where the order parameter

density

is well defined its fluctuations

being typically

of order I

IN

where N is the system size in the case of the mean

spherical model,

the

possible

values run over a wide range. Associated with each

possible

value of the order parameter and the

corresponding

value of the

system

free energy, is a

probability density.

The fact that this

probability density

has a finite

width,

even in the limit N

going

to

infinity,

leads to

nonselfaveraging

behavior and to the

possibility

of

defining

an order parameter function such as the function for the

spin glass

order

parameter Q

in

[5].

In contrast, when the

length

constraint is

strictly

enforced

(as

in [6] where the

spherical

model for the

spin glass

was first

considered, by

Kosterlitz et

al.)

fluctuations are of order I

IN compared

with the average, and the order

parameter

does have a well defined value in the

thermodynamic

limit.

Finally,

to address the

question

of nonrandom

models,

the

property

of

large

fluctuations is shared

by

the uniform

ferromagnet,

which also

undergoes

a form of Bose

condensation,

as noted in

[7]. Notably however,

in this case the

application

of an infinitesimal extemal field has the

property

of

sharpening

the distribution of

magnetization

to a

spike

of

vanishing

width in the

thermodynamic

limit. Hence for the

ferromagnet,

one

expects

that the order parameter is

a well defined

quantity, except

in

strictly

zero extemal field. In the

spin glass

an infinitesimal external field does not have this effect and fluctuations remain

important.

It should be noted that the fluctuations in the mean

spherical

model that lead to the definition of an order parameter function have a different

interpretation

from those described

by

the Parisi solution for the

Ising spin glass [8, 9].

In the case of the

spherical model,

the fluctuations are

entirely

due to the

length

fluctuations

permitted by

the relaxation of the

spherical

constraint. If one were to hold the overall

spin length fixed,

as we remarked

before,

(4)

these fluctuations would be

insignificant. However,

the fact that there is an

underlying disorder,

embodied in the random interactions between

spins

is

crucially important

for the existence of

nonselfaveraging

in the

spherical spin glass.

This is the reason for the difference

in behavior from the

ferromagnet

of Berlin and Kac

[I],

which

(when

considered in the

presence of an infinitesimal external

field,

which can later be set to

zero)

does have the

selfaveraging

property even in the low

temperature phase.

Thus the mean

spherical

model for the

spin glass

is of

particular

interest. It admits the definition of an order

parameter

function that resembles that of the

Ising spin glass,

but unlike the latter owes its existence

only indirectly

to the disorder. One may

obviously object

that the

analysis

here may be far from

applicable

to any real system,

dealing

as it does with a

highly

artificial

(although arguments

may be made in its

favor) length

constraint on the

spins, and,

in

addition,

considers the mean field or infinite range limit

(although that,

as

well,

may be

argued

not to be unreasonable in real

systems). Nevertheless,

there may be valuable

insights

to be

gained given

the

interesting

structure described in this introduction. One great

advantage

of the model is that the fundamental distribution

underlying

the

physics

can be

readily calculated,

and thus it

provides

an

alternative, exactly soluble,

model of a

system

with

large

fluctuations and

nonselfaveraging

behavior.

And,

as we will

show, interesting

time

dependences

in several different

regimes,

such as the stretched

exponential decay

at short times emerge in a natural way.

we

begin

with an outline of the

model,

and then present the main results obtained.

Model for relaxational

dynandcs

in the

Spherical spin glass.

The Hamiltonian of the

Spherical Spin glass

model is H=

-1/2 £ (jS, ),

every

pair

(I, j

of spins

being

assumed to interact with some fixed random

coupling

,,j

(~.

The

spherical length

constraint is that the continuous

spin

variables

($) satisfy

£ S)

=

N. In this infinite range

model,

the

couplings

are all taken from a

single probability

distribution

P[(~]=l/fi~ exp(-N((/2J~).

The different realizations of the

N x N matrix of

couplings J, corresponding

to different

samples,

then form a Gaussian random matrix ensemble whose

eigenvalue spectrum

is known in the limit that N

- co

[10].

Written in

diagonal form,

this Hamiltonian is H=

-1/2 £AS(

where the are the

eigenvalues

of the matrix J. The new

spin

modes

(S~)

are related to the site variables

if by

an

orthogonal

transformation

$

=

£a;~ S~

where the a;~ are functions of the

A

random

couplings,

and are thus time

independent

for a

given sample.

The time evolution of this system, like its

thermodynamic properties,

is easiest treated in the

eigen

basis of J. The low temperature

phase

is characterized

by

a

macroscopically large

amplitude

of the

spin

mode

corresponding

to the

largest eigenvalue

A

= 2J. The critical

temperature corresponds

to that at which the

susceptibility

associated with this mode

diverges,

at T~ = J. In the mean

spherical model,

it can be shown that in zero

field,

the order

parameter is

simply

the

expectation

value

(S( ~j) IN

= q =

(T~ T)/l~ (where

the

angular

brackets denote the thermal

average).

It is this

quantity

that has been demonstrated to possess very

large

fluctuations in the static mean

spherical

model. The strict

spherical model,

on the contrary, does not exhibit

anomalously large

fluctuations in the condensed

phase.

Assuming

a relaxational

dynamics

for the

system,

we

begin by considering

sufficient time to have

elapsed

that initial conditions have

decayed

away

(the

time scale on which this

happens

will be seen

below).

In this case the

equal

time correlations assume their

stationary

values

(5)

662 JOURNAL DE PHYSIQUE I N 5

given by

the effective Hanldltonian

H~~= £(2z- PA )S(

where z is the « chemical

A

potential» conjugate

to the total

spin length.

It is chosen such that the weak

spherical

constraint is

satisfied, namely

£ ($~(t)

=

N

(I)

;

The average denoted

by

the

angular

brackets from this

point

on is to be taken over the random noise

appearing

in the stochastic

spin equation

of motion written below

$(t)

=

ro)

+

1~;(t) (2)

with the Gaussian random noise

obeying (Y~;(t)Y~y(t')) =2ro3;y3(t-t'). ro

is a

microscopic spin flip frequency.

The

simplest (site-,

as well as disorder- in tl~is

case) averaged spin

correlation function in this

asymptotic

time

translationally

invariant

regime

is

C(i)

=

( £ (s;(ii) s,(ii

+

i))

=

£ c~ (i) (3)

where

Ci (t)

= N~

(Si (0) Si (t))

are the correlation functions in the transformed variables.

In the

stationary regime,

the

equal

time

spin expectation

value in

equation (I)

becomes

C

(0)

=

£ Ci (0)

=

£ (2

z

PA )~

=

l

(4)

~

using

the

equation

of motion

(2),

after

transforming

to the

diagonal

basis of J.

Equation (4)

is

just

the

spherical

constraint relation obtained in the static

problem,

and is used to determine

the parameter z as a function of the

temperature

T. The solution for z in the

high temperature phase

is

[6]

2 z

= I +

p ~J~.

In the condensed

phase,

the solution is 2 z

=

2

pJ+ N/q

where

q is the order

parameter

defined

already.

The solution for q represents an average

value, by

which we mean here that the

spin expectation (fj)

can take on

a wide range of

values,

and its distribution is

given by

the Kac

density

function v. The definition of this function and its

explicit

form are

given

in the section on the low temperature

phase.

A note on the time

dependence

of the correlations : the functions

Ci (t)

in this model

decay

to zero at

long

times even in the condensed

phase,

because the

magnetization

remains zero in the absence of any extemal field. This can be seen

by applying

a uniform

field, solving

for the

magnetization subject

to the

constraint,

and

letting

the field tend to zero

ill].

Thus our calculation refers

only

to the condensed

phase

for the

spin

model that is

analogous

to the

superfluid phase

in the Bose gas. The relaxation times and the time

dependence

of the various modes will now be considered in the various

temperature regimes

of interest.

Spin

correlations above the transition.

In the

high temperature phase (Tm J~,

a

straightforward analysis yields

C

(t)

=

j

CA

(t)

=

& ~

dx

fi(f x)

e

~ ~~~°~~ ~~

(5)

using

the

Wigner

semicircular distribution

ii 2]

in

doing

the sum over

eigenvalues (rescaled

to

(6)

lie in

[- I,

I in the second

step (I

=

z/pJ~.

For

sufficiently long

times 2

pJro

t »

I,

one

finds for

example

at

temperatures

much greater than the critical temperature that

C(t)

m

(4 gr)~

~'~

(roJ/Tt)~~'~e~~°~ (Recall

that t is defined as the difference in times of measurements of the

sfins

in the correlation function

equation (3),

whereas the time tj

elapsed

since

letting

the

system

relax

freely

is

always

assumed to be

greater

than any of the characteristic relaxation times of the

system,

which therefore is

independent

of the initial

conditions).

The onsite

spin

correlations thus

decay

with the

expected

rate of

ro

for

temperatures

T» T~.

Close to the transition and

just

above it we may

expand

in

=

(T- T~)/l~,

and for

sufficiently long

times it is found that

C(t)

m

$~(2 grro t)~

~'~ e~~°~°~

(6)

which

gives

the time scale

goveming

fluctuations

r

(T)

= I

fro

~ near the transition. We see that as in the

Ising model,

zv

=

2

(meaning

here

by

z the

dynamical

critical

exponent).

If one

assumes a mean field correlation

length exponent

as in the

Ising model,

this

yields

z = 4 in accord with the

expected

Van Hove result for relaxational

dynamics

for

Ising-like

models

[13].

In

fact, equation (6)

for the

spin

correlation is the same result as found for the

Ising spin glass

model close to the AT line and

just

above it

(see

Ref.

[9],

p.

880].

At

l~ exactly,

in the

spherical

model the correlations

decay

as t~ ~'~ This is in accordance with results from a

variety

of models with disordered

Hamiltonians,

such as the random axis

model

[14],

the SK model

[15],

the random

anisotropy

model

[16]

and in numerical studies of the ± J short range

spin glass [17j.

Thus in the

paramagnetic

as well as critical behavior the

spherical

model

reproduces

some familiar results. Now we turn to the low temperature

phase,

where it is necessary to consider more

carefully

the various averages that are to be

performed.

Spin

correlations in the low

temperature phase.

When T

< T~ it is necessary to consider the contribution of the critical mode

separately

from that of the

remaining

modes. Above T~ the fluctuations in the mean

spherical

model are

negligibly

small and do not

play

any role. However

they

do so for T<

T~,

where the fluctuations in

expectation

value of the condensed mode

6~j

are

important.

In the low

temperature phase,

we have seen,

phases

that are

macroscopically

different contribute to the

thermodynamic properties.

In the low

temperature phase,

the

equation

of motion for the condensed mode contains information

only

about the average relaxation rate of correlations in

Si

2J, which is

ro(2

z 2

pJ).

The

equal

time

expectation

value

(S)j(t))

which is time

independent

is

given accordingly

to be

(2

z 2

pJ)~

=

Nq.

This is the

expected

average

value. However there are

large

fluctuations in this

quantity.

The basic

quantity expressing

the

physics

of the mean

spherical

model is the Kac

density

v

(p T),

which

gives the weight

of the contribution of an ensemble of

spins

with different

spin length normalization, namely

of an ensemble that satisfies the condition

£ ii

=

pN.

The

thermodynamic properties

of the mean

,

spherical

model are

weighted

averages with respect to the measure v, I.e.

Am(T~

=

dp v(p; T) Aj(T)

in which A ~ is a

thermodynamic

observable calculated in the mean

spherical model,

for which the normalization is fixed as in

equation (I),

while A~ is the same observable evaluated

using

(7)

664 JOURNAL DE PHYSIQUE I N 5

the

strictly

constrained

ensemble,

with a value of p not

necessarily equal

to

unity.

This is the Kac relation

[18]

relevant to the

spin problem.

We now take the

point

of view that v

represents

a

probability

of

finding

a

given

canonical ensemble of

spin length density

p # I in the mean

spherical

model

(where

we have set p =

I).

We then

write,

in

effect,

a

generalization

to time

dependent properties

of the Kac

relation as follows

G

(t T)

=

dp

v

(p

;

T)

G

~(t T) (7)

where G

might represent,

for

example,

the

spin

correlation function. The relation above would hold

provided

the system has attained the time translation invariant

asymptotic

state,

and also

providing

that the time t is small

compared

with the characteristic times

r(p )

= I

fro (2 z(p )

2

pJ),

after which the various

p-ensembles

would

begin

to mix. Thus the time scales of observation for which the observations hold is

necessarily

short except rather close to the

transition,

where the characteristic times

diverge.

For

sufficiently

short times

then,

we argue, the evolution of the

spin

mode will be a

weighted

average of the

evolution of states of different

spin length density.

As these have a

macroscopically differing

value of free energy

they

would not be

expected

to be connected

by spin

fluctuations for very short times. At

longer times,

the energy differences may be

bridged,

so that many ensembles

are

averaged

over, and for times

longer

than the system

size,

one would obtain behavior

averaged

over all ensembles and thus described

by

the

single

average time scale

(2

z 2

pJ)~ '.

The

weight corresponding

to a

spin

ensemble whose

length density

» is p is

[5]

)

(p po)

~~~

~

~ ~ ~~ ~°~ ~~~

(x) being

the theta

function,

and with po =

T/J.

It remains now to calculate the correlation function

C(j

for a

given

value of

p in the

strictly

constrained model, and

integrate

with

respect

to the measure in

equation (8).

The calculation of

C(j

is done once

again

with reference to a stochastic

equation

as in

equation (2),

however this

time,

in

place

of the effective

Hamiltonian,

it is necessary to substitute the

appropriate quantity.

This is done

by recognising

that the relaxation rate which is

simply given

in the mean

spherical

model

by ro

3~

H~~/36~

has now to be

computed

for the

case of the

strictly

constrained model. The relaxation time obtained

by looking

at the energy

change

due to fluctuations is a factor N down from that obtained for the mean

spherical

model. This follows from

calculating

the free energy

change

due to small fluctuations about

the mean value

occurring

in the

strictly

constrained

model,

to get

r~'(p)cc

N(2 z(p

2

pJ),

a

quantity

of order one.

Using

this in the

Langevin equation

one finds

(t

<

O(N))

Cjj(t)

=

(p

p

o)

e~ ~'~~ ~°~

(9)

The statistical average over ensembles is defined

by (temperature

T < T~ understood fixed and for t < O

(N))

c~j(t)

=

dp

v

(p) cjj(t)

=

fi(2 qro t)3'4 K~,~( ro t/2 q) (io)

= q

t/r~~

+ i e~

fi

(8)

where an average time

r~~=q/2ro= @/2ro

was introduced

(as

well as the

identity K~,~(x)

=

fie~~

1+

). Thus,

for times

longer

than the time scale r~~ this

yields

x

C~j

~

t~'~exp (-

~),

while for short times it is a stretched

exponential

without the power law

prefactor.

Note

again

that this is

expected

to hold for short times

(times typically

accessible to

experiment)

in a time window which

expands

as one

approaches

T~ from below.

This non

exponential

relaxation

essentially

due to the

averaging

of

dynamical

response over all states

permitted

thus arises in a natural fashion in the mean

spherical

model. At very

long times,

t cc

N,

there is a crossover to the average

exponential decay given by

C~ fit)

- e~ "'~°

(l I)

The full response below T~ contains contributions from the other modes as

well,

C(1)

=

C2J(t)

+

Crest (t)

with

(for

t < O

(N))

~rest

~~~

"

J« ) J)

~~~~

The inverse square root

dependence

appears in the condensed

phase

in other systems such

as the random axis model

[14]

and the random

anisotropy

model

[16],

however in the present model this

dependence

holds

only upto

times of the order of

N,

after which the

decay

becomes

exponential

as in

equation (11).

To

reiterate,

the above low

temperature

formulae are

expected

to hold

only

for times shorter than the time scale

expressed by

r~~. As we remarked

before,

in the

long

time limit

length

fluctuations would get

arbitrarily large.

Then one would use the average relaxation rate in

equation (2)

instead of

performing

an average over ensembles

evolving separately

as in

equation (7).

Discussion.

In

conclusion,

we have

presented

one

possible

formulation for the short time

dynamical

response of the mean

spherical model,

in an extension of its known static

properties.

We Stress here that this

represents

a

possible

but

by

no means

unique

formulation of the

spin dynamics

within the context of a

spherical

constraint. It is based on the

plausible assumption

that the

long

time

dynamics

differs from

dynamics

at short

time,

since the

energetic

cost

renders

large spin

fluctuations

improbable

at short times.

Here we have considered zero field correlations

only.

As we mentioned

earlier, application

of a finite uniform field does not result in any dramatic

change

in the

properties

of the

model,

as far as the

large

fluctuations are concemed.

We comment now on the

applicability

of these results to

experiments

on the

dynamical

response of

spin glasses.

In

spite

of the

oversimplified

character of the

spherical approxi-

mation

itself,

it can be

argued

that the model as

presented

here allows for some intrinsic features of

spin glasses

in a

fairly

natural and tractable fashion. One feature is of course the

non

selfaveraging property

of the

spin

correlations in the

spin glass,

and the second

(related)

is the existence of a

spectrum

of time scales in the system. These

admittedly

have their

origin

in the way that the constraints are

imposed

in the

system,

but it can be

hoped

that their

consequences share

something

in common with the models that have been

proposed

for

random systems, and with the measurements made on real

spin glasses

where it should be

JOURNAL DE PHYSIQUE T I, M 5, MAT 199> 28

(9)

666 JOURNAL DE PHYSIQUE I N 5

recalled the

longest

times of observations are

quite small,

and

only

the very shortest time scales of response are

explored.

The correlations studied here

could,

for

example

be looked for in zero field muon

spin

relaxation

experiments,

which

probe

local

spin

correlations

(a

review of such

experiments

can be found in

[19]).

In the above we have not studied the

interesting dynamical

effects in

spin glasses

that are due to so-called

waiting

time

dependence.

In this paper it is assumed that the initial state is

one in thermal

equilibrium.

Effects due to

switching

on extemal fields and

studying

the relaxation of the

partially equilibrated system

have been studied

by

several groups

[20],

and

are not considered in this paper. Here we note

only

that in

equilibrium,

we

predict

a non

exponential decay

of the

spin

autocorrelations at short

times,

with a

temperature dependent

time scale

given by

r~~ cc @.

Finally,

a comment on the

interpretation

of the spectrum of relaxation processes embodied in

equation (10),

which is a

superposition

of

exponential decays

with different characteristic

decay

times. There exist a number of models for

glassy

relaxation which assume the existence of a distribution of relaxation times. These are often taken to arise due to a distribution of

energy barriers that the

system

is

supposed

to

leap

over. A hierarchical

dynamics

based on activated processes is found for

example

in

[21].

Unlike these

models,

the relaxation times in this paper refer to characteristic times associated with small oscillations around the

multiple

allowed states of the mean

spherical

model. Therefore there is no

Vogel-Fulcher

law for the relaxation time r~~ in the

spin decay.

The short time behavior crosses over

(the

model has

nothing

to say about the crossover

region)

at very

long

times to a

simple exponential decay

with a

macroscopically large

time scale which is associated with

macroscopic

excursions in state space.

Acknowledgment.

I thank J. Rudnick for useful

discussions,

and the Centre de

Physique Thkorique

at Ecole

Polytechnique

for its kind

hospitality.

References

[I]

BERLIN T. H. and KAc M.,

Phys.

Rev. 86

(1952)

821.

[2] READ N. and NEWNS D., J.

Phys.

C 16

(1983)

3273 ; COLEMAN P.,

Phys.

Rev. B 35

(1983)

5072.

[3] ARovAs Daniel P. and AUERBACH Assa,

Phys.

Rev. B 38 (1988) 316 ;

SARKER

Sanjoy,

JAYAPRAKASH C., KRiSHNAMURTHY H. R, and MA Michael,

Phys.

Rev. B 40 (1989) 5028.

[4] ZiFF R. M., UHLENBECK G. E. and KAc M,,

Phys.

Rep. 32C

(1977)

169.

[5] JAGANNATHAN A., RUDNiCK J. and EVA S., to appear in J.

Phys.

A.

j6] KOSTERLiTz J. M., THouLEss D. J. and JONES R. C.,

Phys.

Rev. Left. 36

(1976)

1217.

[7~ GUNTON J. D. and BUCKiNGHAM M. J.,

Phys.

Rev. 166

(1968)

152.

[8] PARisi G., J.

Phys.

A 13

(1980)

l101; ibid. 1887.

[9] BINDER K. and YOUNG A. P., Rev. Mod.

Phys.

58

(1986)

801.

[10]

MEHTA M. L., Random Matrices

(Academic

Press, New York,

1967).

ii I]

A different situation obtains if one assumes that the symmetry

breaking

field in the mean

spherical

model can take on values of order one, I.e. that the

physical applied

field is of the order of

@.

In this

case one would find that the induced « random moment »

(3~j)

is

independent

of the field. The fluctuations in this case are very different from those in the zero field case, and are no

longer

extensive in character. The

analysis presented

here is

inapplicable

in this case.

However this situation is

presumably

irrelevant as it assumes an unreasonable

scaling

of the

physical

field

applied.

(10)

[12] WiGNER E., Proc. 4th Can. Math.

Congr.,

Toronto

(1959) p.154.

[13] ZiPPELiUS A.,

Phys.

Rev. B 29

(1984)

2717.

[14] MA S. K. and RUDNiCK J.,

Phys.

Rev. Lett. 40

(1978)

589.

[15] SOMPOLiNSKY H. and ZiPPELiUS A.,

Phys.

Rev. Lett. 47

(1981)

359

Phys.

Rev. B 25

(1982)

6860.

[16] JAGANNATHAN A., SCHAUB B. and KOSTERLiTz J. M., Nucl.

Phys.

EMS

[FS15] (1986)

324.

[17~ The two dimensional model is considered in BINDER K. and SCHROEDER K.,

Phys.

Rev. B14

(1976)

2142 ;

the three dimensional model in OGiELSKi A.,

Phys.

Rev. B 32

(1985)

7384.

[18] KAc M., UHLENBECK G. and HEMMER P. C., J. Math.

Phys.

4

(1963)

216.

[19] UEMURA Y. J. and YAMAzAKI T.,

Physica

109-l10 B

(1982)

1915.

[20] LUNDGREN L., SVEDLiNDH P., NORDBLAD P. and BECKMAN O.,

Phys.

Rev. Loin. Sl

(1983)

911 ; CHAMBERLiN R. V., MAzURKEViCH G. and ORBACH R., Phys. Rev. Lett. 52

(1984)

867.

[21] PALMER R. G., STEIN D. L., ABRAHAMS E. and ANDERSON P. W.,

Phys.

Rev. Lett. 53

(1984)

958.

Références

Documents relatifs

amplitude modes in the biaxial phase mix the components of biaxiality and birefringence of the order parameter, the critical (quasi-biaxial) mode presenting

2014 We present results concerning sample dependence of not only the modulation wave vector, but also of the amplitude of the order parameter describing the N

Blue Phases I and II, occurring ina small temperature range between an isotropic and a cholesteric phase, have cubic structures with lattice constants of the order of wavelength

In this Appendix, we briefly review the theory of the de SQUID in so far as it concems the effects of trie intrinsic phase shift in YBCO-Pb SQUIDS. We wiii assume that in zero

For evaluating these propagator we define an effective Lagrangian of the fields bQ$~(t,t') (which represent the fluctuations around the MF order parameter value Q$~(t,t')) through

Abstract. This supports the arguments that the long range order in the ground state is reached by a spontaneous symmetry breaking, just as in the classical case...

formidable experimental difficulties (e.g. laser heating) involved in using light scattering experiments to measure the criti- cal exponents a (specific heat) and (order

H. ORDER PARAMETER IN THE GLASS TRANSITION OF VITREOUS S-Ge.. It Was found that a side peak at 440 cm is skronger for the slowly quenched glass than for the rapidly quenched