• Aucun résultat trouvé

Ternary Interactions in Dynamics of Θ-State Polymers: A Conformation-Space Renormalization-Group Approach

N/A
N/A
Protected

Academic year: 2021

Partager "Ternary Interactions in Dynamics of Θ-State Polymers: A Conformation-Space Renormalization-Group Approach"

Copied!
13
0
0

Texte intégral

(1)

HAL Id: jpa-00247178

https://hal.archives-ouvertes.fr/jpa-00247178

Submitted on 1 Jan 1996

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.

Ternary Interactions in Dynamics of Θ-State Polymers:

A Conformation-Space Renormalization-Group Approach

Yasuhiro Shiwa

To cite this version:

Yasuhiro Shiwa. Ternary Interactions in Dynamics of Θ-State Polymers: A Conformation-Space Renormalization-Group Approach. Journal de Physique I, EDP Sciences, 1996, 6 (1), pp.97-108.

�10.1051/jp1:1996131�. �jpa-00247178�

(2)

Ternary Interactions in Dynamics of H-State Polymers:

A Conformation-Space Renormalization-Group Approach

Yasuhiro

Shiwa

Division of Natural

Sciences, Kyoto

Institute of

Technology, Matsugasaki, Sakyo-ku,

Kyoto 606,

Japan

(Received

ii

July

1995, received in final form 19

September1995, accepted

6 October

1995)

Abstract. We reconsider the t'Hooft-Veltman type renormalization treatment of

polymer

properties near the

theta-point.

We correct other recent studies, and now demonstrate that the equivalence between this renormalization scheme and a zero-component field

theory

or a direct

renormalization method for

polymer

chains works

perfectly.

We also

give

an estimate for the

magnitude

of the renormalized third virial coefficient

using

a new result for the translational diffusion constant at the

theta-point.

PACS. 05.20-y Statistical mechanisms.

PACS. 05.60+w

Transport

processes:

theory.

PACS.

61.25Hq

Macromolecular and

polymer solutionsj polymer

melts.

1. Introduction

When the

temperature

of a

polymer system

of

long

flexible chains diluted in a

good

solvent is

lowered,

the

repulsive binary

interaction between

segments

on the chain decreases in

strength

and becomes

attractive,

while the

ternary

interaction becomes

relatively

more

important.

At the so-called

fl-temperature,

the two- and the

three-body

interactions

balance,

and the second

virial coefficient vanishes.

It is

generally accepted

that the

fl-point

is a tricritical

point

with upper critical dimension

dc

= 3

[ii.

This

implies

that the critical behavior in three dimensions is of mean-field

type

except for

logarithmic

corrections

[2].

These corrections come from the

three-body

interactions.

However,

the

precise

form of the

logarithmic

corrections has not been without

dispute.

Kholodenko and Freed [3] criticized that the field theoretic calculations [2]

using

the

equivalence

between the

polymer system

and the

O(n)-#~

model in the limit n - o (~§ is an J1-component

field)

gave wrong results.

Later,

Freed and his coworkers [4] also claimed that dimensional

regularization

calculations for the continuous

three-parameter

model [5] of the

polymer

would be incorrect in three dimensions and that the cut-off method must be used

[6].

Both these

points

were refuted

lucidly

in a series of work

by Duplantier ii, 8].

He showed that the J1 - o

#~ theory

and the

polymer three-parameter theory

gave

exactly

the same results. He also elucidated the relation between the cut-off and the dimensional

regularization

schemes to demonstrate the

validity

of the use of dimensional

regularization.

The above

controversy

was

mainly

concerned over static

properties

of

polymer solutions,

and the

logarithmic

correction in the

dynamic properties

has not received that much attention. The aim of the

present

paper is to

present

the first correction to the

mean-field

behavior of the

©

Les

(ditions

de Physique 1996

(3)

translational diffusion constant at the

fl-point.

For

this,

we shall

rely

on the

three-parameter

model of

polymer

chains with the

help

of the t'Hooft-Veltman

style

renormalization scheme

(to

be referred

collectively

as

conformation-space renormalization-group

method

[9]). Actually.

the

problem

was considered

already by

Freed et al. [4] in the same framework.

Holn-ever,

the entire calculation is

plagued

with the difficulties described above.

Moreover,

we note that the calculations with use of the

conformation-space renormalization-group

method have never been

presented

for

polymer properties

in the

fl-region during

the

dispute (although

the so- called direct renormalization method of des Cloizeaux

[10]

was

applied

to the

three-parameter model,

and the relation between the direct and the dimensional renormalization schemes ~vas studied

[7]).

In view of this

fact,

it is desirable to users of other formulations to recover the

previous

results

directly by

the

conformation-space renormalization-group

method.

This,

too.

~vill be demonstrated in this paper.

The paper is

organized

as follows. In the next section the

conformation-space

renormalization- group formalism for the continuous

polymer

chains near the

fl-point

is

given.

We calculate the

leading logarithmic

correction to the

asymptotic

law for the mean end-to-end distance as well

as the mean radius of

gyration

of a

polymer

chain. The results coincide

precisely

,,~ith the

previous

results. In Section

3,

the

foregoing

formulation for the static

polymer properties

is extended to treat

dynamics.

Based on the kinetic model

[11]

of

polymer solutions,

we calcu- late the translational diffusion constant at the

fl-point. Computational

details are

pro,~ided

in

Appendix.

The paper concludes with a discussion in Section 4.

2. The Model and Renormalization

2. I. THREE-PARAMETER CoNT~Nuous-CHA~N l/IODEL. We consider a

single

flexible

poly-

mer chain immersed in a solvent in 3 dimensions with a conformational

probability given by

[5]

P(r)

=

exp[-H(r)],

No

d~j~)

b No No

Hlrl

=

j / dsl~l~

+

j /

dS

/ ds'blrls) rls'))

o o o

~ No No No

+ ° ds

ds' ds"b(r(s) r(s'))b(r(s') r(s")) ii)

6

Here

r(s)

denotes the

position

of the chain

parameterized by

the contour variable s

lo

< 5 <

No No being

the contour

length(~

of the

chain).

Coefficients bo and co are the two- and the three-

body

interaction

parameters, respectively.

One may have an

arbitrary sign

for

bo,

while co is

positive

for

stability.

In this so-called

three-parameter

model we have two dimensionless interaction parameters.

They

are defined

by

zo +

(21r)~3/~boNll~

, yo +

(21r)~~co (2)

2.2. RENORMAL~zAT~ON ~N THE fl-REGmN. The model

(I)

is ill-defined since there are

ultraviolet

divergences appearing

in the

perturbation expansion

of some

macroscopic

quan-

tity

calculated from it. One then has to eliminate these

singularities.

This can be done

by

renormalization

techniques,

and we use diniensional

regularization [12].

(~) In

fact,

the parameter No has the dimension of an area c~

(length)~.

This

originates

from the fact that the fractal dimension of an ideal chain is two. In this paper,

however,

we use the convention of the conformation-space renormalization-group theory to call No the total

length

of the chain.

(4)

Let L be a

phenomenological length

scale,vhich has the same dimension as )To We may then define the dimensionless

coupling

constants ito and MO in d space:

uo %

(21r)~~/~b,nL~/~

,

wo +

(21r)~~coR

,

(3)

where e e 3

d,

and we have introduced a

marginal two-body

interaction

parameter

[7]

bm e

boNll~

,

(4)

which is

marginal

for d

= 3 but not for e

#

o.

Let us consider the mean square end-to-end distance

R2

and the

two-body

osmotic pa-

rameter g of an isolated

chain;

in terms of g, the second virial coefficient

A2

is

given by

A2

"

Ii /2)g (21r-~2)~/2,

~vhere

X2

=

R2 Id.

Their

expansion

in po~vers of ~o, MO and e may

be evaluated in the

straightforward

manner

(cf.

Ref. [13]

).

The results read

~~ ~'~~ ~~~°°

~

~~~~°~

~

~°~~ ~~~°°~

~' ~~~

g =

-8wo Ii ~~~wo)

+ ~o

Ii ~~~wo)

+.

(6)

e e

The

ellipsis

denotes third- and

higher-order (in

uo and

wo)

terms. It should be remarked that in the above we have

kept only

the dominant contribution in e to each term.

We assume that the model

(1)

is renormalizable with the

follo,ving

renormalization

factors,

the

Z's,

and introduce the renormalized

quantities N,

iu and u:

No"Zp~iv, wo=Z~vw, uo"Z~u. (7)

Since

singularities

arise from the

three-body interactions,

the Z's

depend only

upon w. Hence

,ve introduce the

expansions

Zi~

"

1+Aw+Biu~+.

,

Z~

= I+Ciu+.

,

Z~

= + Diu +

(8)

Because

R2

and g are

directly observable, they

should not be renormalized. Then we

readily

find that

~

A

= o

,

B

=

~~~

,

C =

~~~

,

D

=

~~~

(9)

3e e e

zw

= 1+

~~~

aJ +

o(w2) (ii)

Z~

=

1+

~~~ MJ + (MJ~) . (12)

(Notice that (10) orrects result of

Ref.

the

lvilson

unctions

[12]

~ to

(5)

the

derivative 3/3L being

taken

with

the

fixed

renormalized

parameters

No

and

Now,

define

the

w(L)

L

~

AT =

Ni(MJ) (16)

Equivalently, introducing

the dilation factor t with

dL/L

= dt

It

[9], we define

w(t)

and

-V(t) by

The

solution

2.3. AT THE fl-PO~NT IN 3-SPACE. In this

subsection,

we consider a

system

for which

u = o

(at

the

fl-point)

and d

=

3,

and calculate

R~

for

large

N. lye

shall,

of course, obtain the

logarithmic

correction to the mean-field

(ideal-chain)

form.

Going

to the limit e - 0 in

(17),

we obtain the

equation t~~°

=

-441riu~

,

(19)

which is

easily integrated

to

yield

~°~~~

1 +

~j~~1)

In t ~~~~

Combining (18)

and

(20),

we then find

~~~~ ~~~~ ~~~~~~ +~~~j~~~)In

t~ ~~~~

The next

step

is to

employ

the

matching procedure [15]. Namely,

we continue the solutions

(20)

and

(21)

until the

perturbation expansion

for

R~

can be

trusted;

we shall choose t

= to

at which

iu(to)

<1.

l~Thereupon

the loin-est order

perturbation theory is) gives R~

=

3N(to)(1 41rai(to)) (2?)

With the initial conditions

Nil)

=

N,

MJ(1)

= y

corresponding

to the

physical

chain of

length

N with a

coupling

constant y, we choose to

"

lk~la,

where a is a small cut-off constant.

Then,

writing

h % MJ(to) so as to make contact with the notation of reference

[13],

we

finally

obtain from

(22)

~~ ~~~~°~~~~~ ~~~~~

'

~°~~~

~

~~~

~~~~

(6)

with

~

41rlu(Nla)

~~~~

In the literature h is sometimes referred to as the renormalized third virial coefficient. We note in

passing

that the condition of

validity

of

(24)

is y

In(Nla)

»

1/441r.

Similarly,

from the lowest-order

perturbation

result for the mean square radius of

gyration:

R[

=

~N(to)(i ~~~iu(to))

,

(25)

~~~~

(26)

we

get

~2

=

NAO(l/)(1 W~~

Expressions (23)-(26)

agree with the results obtained in references

[7,13]

via different methods of renormalization.

2.4. 8-REGmN ~N 3-SPACE. We now assume that zo is small but finite in order to

study

the

fl-region.

In this case, too, we can follo~v a similar sequence of steps as in the

preceding

subsection.

Moreover,

it is necessary to consider the renormalization flow

equation

of ~ beside

(17)

and

(18).

We first find

~/L~~~°'~°'~"~

(~ ~~~~"~'

~~ +

Ll~

in z~)N~,co,b,~~ = 167rw +

OIW~ 127)

In

obtaining (27),

let us recall that it is essential for the L-derivative to be

performed

at fixed bm

iii. Thus,

for d

= 3 and to the lowest

order,

t~~

= -161r~tiu

(28)

This is

integrated

with use of

(20)

to

yield

~llt)

=

Ji41v)zlh/v)~/~~

,

129)

where we have set ~c(1) =

Ji4 IV)?(

z is the renormalized

counterpart

of zo and

Ji4

IV) is some

as

yet unspecified

function of y.

From the

perturbation expansion

R~

=

3N(to)(1 41riu(to)

+

~~c(to))

,

(30)

we then obtain

R~

=

3NjAo(Y)11 ~~~h)

+

ZA41Y)l~)~~~~l

131)

with

A4(y) +1i4(y)-40(Y). Furthermore,

the radius of

gyration

can be deduced from

to be

~~ ~~~°~~~~~ ~~i~~

~

~~~~~~~~~~~~~~~~

~~~~

(7)

Comparing (31)

and

(33)

~vith the first terms of the well-known

expansions R2

=

3N(1+~z+. ),

RI

=

[Nil

+

fz

+ ~~~~

which are calculated for a chain with

only two-body

interactions

(see,

for

instance,

Ref.

[13]),

we find

A4(o)

=1.

(35)

Consequently,

as advertised in the

introduction,

the results

(31)

and

(33)

are in exact

agreement

with those obtained

by Duplantier iii

and des Cloizeaux [13]

through

the

zero-component

field

theory

or the direct renormalization method.

3. 3lranslational Dilfusion Constant at the fl-Point

Having

formulated the

conformation-space renormalization-group theory

of static

properties,

,ve now

proceed

to examine

dynamics.

In

particular,

we shall focus on the translational diffusion at the

fl-point.

~Ve use the usual set of

Langevin equations describing coupled

chain-solvent

dynamics ii Ii

The solvent

elocity

v(x, t) at a point ix, t)

is

ssumed to satisfy the

pressibility ndition, T7 . v =

o.

As a consequence, the

tensor

erator T lects

the trans,,erse

mponents

of the

field it

jf~js,t)f~jsl,tI)j

=

2<piiijs sI)ijt tI)

,

ifix,t)fixl,tI))

=

~2~~iv2ijx xI)ijt tI) j38)

being

the unit tensor. The parameters

(o

and ij~ ,vhich appear in our model are the bare friction coefficient per

segment

of the chain, and the local sol,~ent

viscosity, respectively.

The ratio

(o/q~

becomes the relevant

dynamic

parameter in the model.

Based on the kinetic model

(36)-(38),

we can deri,~e the formula for the translational diffusion

constant D of the

polymer

chain. This is achieved

by

the elimination of the solvent

degrees

of freedom in favor of an Oseen tensor

description

of the

polymer [16, Iii.

The result has the form of the Ilirk,vood formula:

D

=

£ ii

+

(jj~Tr Ids / ds'iTiris) ris'))11

1391

Here kB is the Boltzmann cunstant, and

T(x)

is the Oseen tensor defined

by (in d-space) T(x)

=

/ ii kk/k~)e~~'~ (40)

k k

(8)

with

f~

e

(21r)~~ f

dk. We remark that the above formula

(39)

is reliable up to

(~) O((o/~le),

and the average

(.

may be taken with

respect

to the full Hamiltonian H of

ii ).

In the

following,

we calculate the diffusion constant at the

fl-point,

hence we may

put

bo " o

in

ii

from the outset. This is

permissible

if one

performs (renormalization)

calculations

using

dimensional

regularization;

the dimensional

regularization just

measures bo with

respect

to its

fl-point

value

[8].

One may use

AppendiK

A to calculate the

right-hand

side of

(39).

One finds D =

~~~ [1+ ~~(o(1 Aiiuo))

,

No (o

9

(41)

to lowest order in iuo, where

Ai

" o.89582. We have introduced a dimensionless

(in d-space)

parameter pertaining

to our model:

IO "

12~l~~/~fmv/~

,

fm *1(0/~/e)N/~~ 142)

Here we renormalize the

expression (41) by extending

the program

explained

in the

preceding

section. The renormalization factor

Zj

defined

by

lo

"

Z/~( 143)

may be determined with use of

(lo)

and

ill ). (Notice

that since D

depends

on

No

and MO, the renormalization of

lo

is affected

by

the renormalized static effects. The result

is,

to

leading

order in iu and

(,

~

Zj

= 1-

~~

iu~ (44)

e

We can now

proceed

as before. The renormalization flow

equation

for

(

becomes at d

= 3

t~~

=

(L(

~ ln

Zj)No,co,j,»

"

-~~~~~iu~ (45)

Notice that the

logarithmic

derivative is taken at fixed

(m

since

(m

is

marginal

in three dimen- sions

(just

as the

marginal two-body

interaction

parameter bm;

see

(27) ).

The solution to the

equation (45)

is

given by

~~~~

~~~~ ~~~

1

~~~j~In

t~ ' ~~~~

where we have written

(

e

((1).

We observe at this

juncture

that in the limit N - oo,

lit)

-

fexpl-8irv/33) 147)

Since

(

may take on an

arbitrary Ii-e-, sample-dependent) value,

we

have,

at the

fl-point,

a

partial draining

of

varying degrees

even in the

long-chain

limit

(the scaling limit).

This is to be contrasted with the result of the

renormalization-group

calculation which

employs

the usual

die

4

d) expansion.

The

f-expansion theory predicts

[9] that in the

scaling

limit one

always

ends up with the

fixed-point

value for the

draining parameter ((o/~le)L"2

(~)

For

long

chains in

good solvents, (o/~e

may be taken to be of order @(+4

d).

It is for this reason that in reference [9] it was concluded that there is no

justification

for the Kirkwood formula

beyond

order @. It should be

emphasized

that it incurs

no restriction upon the contribution from the static effects

(such

as the excluded-volume interaction, whose strength is also of the order @).

(9)

Finally,

from

(41)

and

(46)

with the choice

to

"

Nla

as

before,

we obtain the

expression

for D at the

fl-point.

It reads

~)~ ~~D

=

~~ (1+ ~~([Ao(y)]~~/~(1+ Bih))

,

(48)

~~

N

(

9

where

Bi

" 81r

/33 -41.

The second

equation

of

(42) implies

that when the chain becomes very

long,

the second term on the

right

of

(48)

becomes dominant.

Thus, defining

the

hydrodynamic

radius

RH by

the relation

D=

~~~

,

61r~~RH (49)

we

find,

for

large

value of

N,

Rj~

=

~

[NAO(y))~~/~(1+ Bih) (So)

7r

Equations (48)

and

(So)

constitute the main results of this paper.

4.

Concluding

Remarks

An estimate of the

magnitude

of the

three-body (osmotic) parameter

h is now

possible by comparing

the theoretical results of Section 3 with data from

experiment

and simulations. The

quantity RG/RH

is of interest in this respect. It follows from our results

(23)

and

(So)

that the ratio at the

fl-point

is

given by

There are difficulties in the

comparison

of

(51)

with

experiments. First,

the

steady-state

diffusion constant should be measured to determine

RH

as defined

by (49). However, usually

the values of the

hydrodynamic

radius are estimated from the initial

decay-rate

of the

dynamic

structure factor

[18].

The so-called initial diffusion constant thus obtained can be different from our D in

principle.

Another reason for the

difficulty

is that the

precise

determination of

RH

from

(49) requires

an accurate value for the ratio

~~/~~

at the same time

[19],

where ~~ is the bulk solvent

viscosity.

This is at

present

a serious obstacle in

practically comparing

the

theoretical

prediction

with

experiments.

In reference

[20],

Bruns

performed

Monte Carlo simulations to

directly

calculate the

equilib-

rium ensemble average of the second term on the

right-hand

side of

(39). Therefore,

his estimate of

RH

is free from any such

complication

as described in the

preceding paragraph.

From the

value

RG/RH

= 1.464

given

in this reference

together

with

(51),

we estimate h

= 4.5 x

10~3

We can

separately get

an estimate from

R~ /R(

that has been measured in the same simulation.

Comparing

the data

R~/R(

= 5.94 with our

prediction (see (23)

and

(26)

~

=

6(1

~~

h)

,

~ (52)

4

we find h

= 4.2 x

10~~, comparable

with the

dynamical

estimate

given

above. Our values agree

reasonably

well with the value h

=

(1.2

+

o.2)

x

10~3

which was estimated

[21]

in terms of other static measurement in real

experiments [22].

The above

comparison provides

an evidence of the existence of

ternary

interactions near the

fl-temperature.

We

emphasize, however,

that

although

we have a

good agreement

in order of

(10)

magnitude

of

h,

this does not mean that the

logarithmic

corrections to ideal-chain behavior as

predicted by

the

theory

have been verified

by experiments

and simulations. In

fact,

a recent

large-scale

simulation

[23]

with

supposedly high

accuracy has found

logarithmic

corrections which are different from the renormalization theoretical

predictions.

Now that we have several

different theoretical

approaches

to the

study

of the

fl-region,

which are

entirely equivalent,

and since this paper has

provided

the theoretical

expression

for the

dynamic properties,

we

challenge experimenters

to test these

predictions. Meanwhile,

the renormalization theoretical

calculation of other

transport coefficients,

such as intrinsic

viscosity

and

decay

rate, would be very useful. In

brief,

it is

hoped

that the renormalization

method,

which has been so

successful in

elucidating

the solution

properties

of

polymers

in a

good solvent,

may resolve those

problems [24]

that confront the current

theory

of

fl-regions.

Appendix

A

This

appendix

summarizes the calculation of the

right-hand

term of

(39).

We will start with

Tr

/~

ds

/~ ds'jTjrjs) rjs'))j

=

id ljs~ / Hjkj /k~

,

jA.I)

where

f~

e

(21r)~~ f

dk and

H[k]

=

j / ds' /

ds"(e~~'l~~~'~~~~~")I

e

j / ds' / ds"h[k;

,

s', s"] (A.2) Throughout

the

Appendix

let S denote

No

for the sake of notational

simplicity.

The function

h[k; s', s"]

can be written as

[25]

, ,,

Z(k;S,s',s")

jA.3)

~'~'

~

'

~

Zi IS)

Here

Z(k; S, s', s")

is a

partition

function of a chain with two insertions of wave vectors k and -k at contour

positions

s' and s"

along

the

chain,

and hence

Zi IS)

e

Z(0; S,s',s")

is the

single-chain partition

function.

Diagrammatic perturbation theory

to calculate Z is well established

[13].

The

diagrams contributing

Z to first order in co are indicated in

Figure

1. As a

representative example,

we

illustrate the calculation of the contribution of the

diagram if)

to the

integral Z(S)

e ds'

ds"Z(k; S, s', s") /k~

which appears in the

expression

for

H[k]. Labelling

the

diagram

as shown in

Figure 2,

we find that it is

given by

S s"

s3 s2 si

Z~(S)

%

(-C0) / / / / dS" / dS3 / dS2 / dsl / dS'

k k, k2 0 0 0 0 0

k~~G(0,

S

s")G(k,

s"

s3)G(k k2,

s3 s2

x

G(k ki,

s2 si

)G(k,

si

s')G(0, s') (A.4)

with

G(q,s)

=

exp(-q~s/2) (A.5)

(11)

s

.,, .,,

,~~ ,~_

s'

'~

~j j ~ )

~

s' ,;~

"'

'''

~ ~

i

(1) (a) (b) (c) (d) (e) (f)

Fig. 1. The diagrams

contributing

to

Z(k;

S,

s', s")

to first order in three-body interactions. Each

diagram

consists of

one

polymer

line

(solid) representing

a chain of

lenght S,

and of the

three-body

interaction line

(dashed) connecting

the interaction

points

of the chain. Wave vectors k and -k are

injected

at the contour

points

s' and s".

>. s'

S~ '~..

j

k~

S~ 'S)

k~

S, '""'

.>.. s'

Fig.

2. A

diagram contributing

to

Zf(S).

Taking

the

Laplace

transform of

Zf(S)

in

S-space yields

)lip)

=

(-co)P~~ / / / k~~ [©(k, p)]~©(k ki, p)©(k k2,

P)

,

(A.6)

k k, k~

where p is the

conjugate

variable to

S,

and

©iq,P)

=

iq~/2

+

P)~~ IA-I)

No~v the

integral

in

(A.6)

is

performed by

the method of dimensional

regularizatioii. Using

(12)

standard formulas:

Ii Tin d/2)

q [q2 +

2(k q)

+

m2]" 2"lrd/2T(o)(m2 k2)°~d/2

'

~ )~)~(~~ ~ ~[ai~~~~(1 ~~)]~+P

'

~~'~~

we find

jij~)

=

j_~~)~3d/2-7j~~)-3d/2 7rld 4)irli d/2)l~

j

~ g~

21d 2) sin17r13 d/2)lrld/2 1) Performing

the inverse

Laplace

transform of

(A.9)

and

by analytic

continuation to d

=

3,

we

then obtain

js/~)3/2

~~~~~~~

~°~

3~7/2 ~~'~~~

Contributions ofother

diagrams

are evaluated in the same manner. Then

Z(S)

can be

expressed

as a sum of the contributions associated with the

diagrams (o)-(f)

of

Figure

1:

1

zjs)

=

~j g~zijs)

,

jA.ii)

imo

1n.here the

prefactor

gi is the

degeneracy (weight)

of the

diagram Ii),

and

zo

~

4j

S

)3/2

3 21r Z~

=

-41rMJoZ°, Z~

=

-(Z~

,

Z~ =

Z~

=

In(3/2)

x Z~

,

Z~

= ln2 x Z~

,

Zf

=

jZ~ (A.12)

Moreover,

for the

single-chain partition

function we have

Zi(S) s2

=

-(1- 41riuo) (A.13)

Substituting (A.11)

and

(A.13)

into

(A.1),1n-e finally

obtain

Tr

Ids / ds'(T(r(s) r(s')))

=

~~(

~

)3/2 ~~~~~°°

,

(A.14)

3 21r 1

4~wo

where

Ci

= 2 +

In(32/81). Keeping

the dominant contributions

brings

us to the

expression (39)

with

Ai

=

4~(Ci -1).

(13)

References

ill

de Gennes

P.G.,

J.

Phys.

Lett. France 36

(1975)

L-55.

[2]

Duplantier B.,

J.

Phys.

France 43

(1982)

991.

[3] Kholodenko A. L. and Freed K.

F.,

J. Chem.

Phys.

80

(1984)

900.

[4]

Cherayil

B.

J., Douglas

J. F. and Freed I<.

F.,

Macromoiec~ties 18

(1985)

821 and J. Chem.

Phys.

83

(1985)

5293.

[5] Edwards S.

F.,

Proc.

Phys.

Sac. 88

(1966)

265.

[6] We notice that this conclusion was

repeated

in Freed's book: Freed K.

F.,

Renormalization

Group Theory

of

Macromolecules, (John-Wiley

&

Sons,

New

York, 1987) Chapter

11.

[7]

Duplantier B., E~trophys.

Lett. 1

(1986)

491, and J.

Phys.

France 47

(1986)

745.

[8]

Duplantier B.,

J. Chem.

Phys.

86

(1987)

4233.

[9] Oono

Y.,

Adv. Chem.

Phys.

61

(1985)

301.

[10]

Des Cloizeaux

J.,

J.

Phys.

France 42

(1981)

635.

[11]

Shiwa Y. and Oono Y..

Phystca

A 174

(1991)

223.

[12] See,

for

instance,

Amit D.

J.,

Field

Theory,

the Renormalization

Group,

and Critical

Phenomena, (World Scientific, Singapore, 1984).

[13]

Des Cloizeaux J. and Jannink

G., Polymers

in Solution: their

Modelling

and

Structure, (Oxford

Univ.

Press, Oxford, 1990).

[14]

The

three-body parameter

of reference

[3],

WKF, is related to our w

by

WKF

" 2~w.

[15]

Shiwa Y. and Ka~vasaki

K.,

J.

Phys.

C15

(1982)

5345 and reference

[ii

for corrections.

[16]

Lee

A.,

Baldwin P. R. and Oono

Y., Phys.

Rev. A 30

(1984) 968;

see also Baldwin P. R.

and Helfand

E., Phys.

Rev. E 41

(1990)

6772.

[17] Oono

Y.,

AIP Conf. Proc. 137

(1985)

187.

[18] Tsunashima

Y.,

Hirata

M.,

Nemoto

N., Kajiwara

K. and Kurata

M.,

Macromoiec~ties 20

(1987)

2862 and references therein.

[19]

Shiwa Y. and Tsunashima

Y., Physica

A 197

(1993)

47.

[20]

Bruns

W.,

Macromolecules 17

(1984)

2826.

[21] Duplantier

B. and Jannink

G., Phys.

Rev. Lett. 70

(1993)

3174.

[22]

Duplantier

B~ Jannink G. and Des Cloizeaux

J., Phys.

Rev. Lett. 56

(1986) 2080;

Boothroyd A.T.,

Reniiie

A-R-, Boothroyd

C.B. and Fetters

L-J-,

ibid. 69

(1992)

426.

[23] Grassberger

P. and

Hegger R.,

J. Chem.

Phys.

102

(1995)

6881.

[24] Fujita H., Polymer

Solutions

(Elsevier, Amsterdam, 1990);

Shiwa

Y.,

Macromolecules 28

(1995)

1326.

[25]

Duplantier B.,

J.

Phys.

France 47

(1986)

1633.

Références

Documents relatifs

From the recursion equation (9) one obtains also a n attractive fixed point with one IR unstable direction, which describes the critical un- binding transition in lamellar

The critical temperature of the square lattice Ising ferromagnet was first obtained [9] exactly from the.. self duality property of the Ising model on

The renormalization group approach to critical dynamics near the nematic- smectic-A phase transition in liquid crystals... The renormalization group approach to critical dynamics

Second, the renormalized coupling among different polymer species is found to be bounded from above which leads to the existence of a well defined limit of strong

Powerful numerical techniques, like Monte Carlo simulation [2], exact enumerations analysis [3, 4], or phenomenological renormalization group (RG) cal-.. culations on

2014 The scaling of the failure strength as a function of the size L and the influence of disorder in hierarchical and Euclidian systems are discussed within a real

out compared to the swollen Gaussian approxi- mation, so that the chain can decrease its central segment density. Full or broken lines- denote the excluded volume

More complex interactions are generated upon renormaliz~ition, and will be given below... 3, The Triangle-Dart tiling. ai i piece of the iiling : bj ihe eighi type, ut ,ire, wiih